14.06.2013 Views

PDF, about 5Mb - Teksid Aluminum

PDF, about 5Mb - Teksid Aluminum

PDF, about 5Mb - Teksid Aluminum

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

THE JOURNAL IS ALSO<br />

AVAILABLE ON<br />

TEKSID ALUMINUM<br />

WEB SITE:<br />

WWW.TEKSIDALUMINUM.COM<br />

Editorial Panel<br />

Paolo ANTONA<br />

Materials Consultant.<br />

Pietro APPENDINO<br />

Professor, Material Technology and Applied Chemistry; Faculty of Engineering,<br />

Turin Polytechnic.<br />

Marcello BADIALI<br />

R&D Director, <strong>Teksid</strong> <strong>Aluminum</strong>.<br />

Giuseppe CAGLIOTI<br />

Professor, Solid State Physics ; Faculty of Engineering, Milan Polytechnic.<br />

Enrico EVANGELISTA<br />

Professor, Metallurgy; Faculty of Engineering, Ancona Polytechnic.<br />

Luca Paolo FERRONI<br />

Marketing Director, <strong>Teksid</strong> <strong>Aluminum</strong>.<br />

Merton C. FLEMINGS<br />

Head Department of Materials Science and Engineering, M.I.T., Cambridge, Mass.<br />

Sergio GALLO<br />

Invited Professor, Applied Chemistry & Metallurgy; Faculty of Engineering, Turin<br />

Polytechnic. Past President <strong>Teksid</strong> S.p.A.<br />

Cinzia MODENA<br />

Editorial Coordinator, Metallurgical Science & Technology, <strong>Teksid</strong> <strong>Aluminum</strong>.<br />

Claudio MUS<br />

Materials Consultant.<br />

Walter NICODEMI<br />

Professor, Siderurgy & Siderurgical Technology; Faculty of Engineering, Milan Polytechnic.<br />

Editor in chief: Marcello BADIALI<br />

Registrazione presso il Tribunale di Torino n. 3298 del 12 maggio 1983<br />

Associato alla Unione Stampa Periodica Italiana<br />

Authors are asked to supply two typewritten copies of their papers. These<br />

should be laid out in accordance with the “Instructions for authors” on the<br />

inside back cover. All correspondence should be addressed to:<br />

Segreteria di redazione<br />

Metallurgical Science and Technology<br />

<strong>Teksid</strong> <strong>Aluminum</strong> S.r.l. Via Umberto II, 5<br />

10022 Carmagnola (TO) Italy - Tel. +39.011.9794606<br />

e-mail: journal@teksidaluminum.com<br />

Publication distributed free of charge<br />

© Copyright 1983 <strong>Teksid</strong> S.p.A. - All rights reserved.<br />

Editorial coordination: Cinzia Modena, Marketing, <strong>Teksid</strong> <strong>Aluminum</strong><br />

Graphics and layout: EMMEDI pencil & mouse - Via S. Ambrogio, 23 - Torino<br />

Printed in Italy: Graficat - Via Cuniberti, 47 - Torino<br />

No part of the texts published in this journal may be reproduced, whether in<br />

the original language or in translation, without permission in writing from<br />

<strong>Teksid</strong> <strong>Aluminum</strong>.<br />

Papers submitted for publication as a rule should illustrate unpublished original<br />

research works, experiments, or critical reviews. Publication will depend on<br />

the approval of the Editorial Panel.


E. Gariboldi, D. Ripamonti,<br />

L. Signorelli, G. Vimercati,<br />

F. Casaro<br />

S. Seifeddine, T. Sjögren,<br />

I. L. Svensson<br />

M. El Mehtedi, L. Balloni,<br />

S. Spigarelli, E. Evangelista,<br />

G. Rosen, B.H. Lee, C.S. Lee<br />

Metallurgical<br />

Science and Technology<br />

A journal published<br />

by <strong>Teksid</strong> <strong>Aluminum</strong><br />

twice a year<br />

Vol. 25 No. 1, July 2007<br />

FRACTURE TOUGHNESS AND<br />

MICROSTRUCTURE IN AA 2XXX<br />

ALUMINIUM ALLOYS<br />

VARIATIONS IN MICROSTRUCTURE AND<br />

MECHANICAL PROPERTIES OF CAST<br />

ALUMINIUM EN AC 43100 ALLOY<br />

COMPARATIVE STUDY OF HIGH<br />

TEMPERATURE WORKABILITY OF ZM21<br />

AND AZ31 MAGNESIUM ALLOYS<br />

PAGE<br />

3<br />

12<br />

23


EDITORIAL<br />

The present issue marks an important goal for <strong>Teksid</strong> <strong>Aluminum</strong>: the 25th<br />

anniversary of Metallurgical Science and Technology (MS&T). Twenty five years<br />

of life represent a remarkable milestone which is not commonly reached<br />

by a scientific journal. MS&T is one of the few specialized editorial products<br />

edited by an industrial company. It has been always delivered world wide,<br />

free of charge and with continuity.<br />

This moment deserves a word on the nature of the Journal, on its intrinsic,<br />

permanent features as well as on its evolution and future perspectives.<br />

The birth of MS&T originated from the initiative of professor Sergio Gallo,<br />

former President of <strong>Teksid</strong> S.p.A., together with full professors Aurelio<br />

Burdese of the Politecnico di Torino and Walter Nicodemi of the Politecnico<br />

di Milano. They suggested to earmark part of the financial resources destined<br />

to the image promotion activities of the company for a specialized scientific<br />

review. Their proposal was warmly agreed by Antonio Mosconi, the then<br />

Managing Director of <strong>Teksid</strong>, who, in the first issue’s introduction, wrote:<br />

“The aim of Metallurgical Science and Technology is to initiate a dialogue,<br />

already part of the scene in other countries, between all those<br />

resources that are committed to the search for new frontiers in<br />

metallurgical techniques, so as to derive the greatest synergies from<br />

them.<br />

Metallurgical Science and Technology then enters the scene as an<br />

instrument whereby the dialogue between metallurgical science and<br />

metalworking industry can be taken to greater depths.<br />

‘Academic’ metallurgy indeed, has often shown signs of hankering<br />

after pure science, estranged from industry. It should not be forgotten<br />

that its final objective remains practical application.<br />

A journal such as this is not as ambitious as to imaging that it can<br />

entirely fill the goal. Its aim is rather to inculcate the habit of<br />

comparison and so cast a permanent bridge between the two parties.<br />

Its very existence shows that the need to set up a more systematic<br />

and continuous two-way flow of information is seen by both sides as<br />

an indispensable prelude to its development.”<br />

Since these targets and perspectives were indicated, several changes have<br />

occurred in the life of <strong>Teksid</strong> Group, the sponsor and editor of the journal.<br />

The original name of the company, <strong>Teksid</strong>, evokes a core business centered<br />

on cast iron foundry. Founded in 1978 as a spin-off of Fiat metallurgical<br />

activities, its experience rooted in the very beginning of the Italian industrial<br />

development which started-up in the early century. Steel and cast-iron<br />

production were then the company’s core business focus. The global trend<br />

toward “energy conservation” triggered by the recurrent oil crises of the<br />

seventies and eighties induced <strong>Teksid</strong> to share and favor the efforts aiming<br />

at vehicle weight reduction: the interest of the Company then gradually<br />

Metallurgical Science and Technology<br />

shifted to light metal alloys – like magnesium and,<br />

above all, aluminum – as base materials. <strong>Teksid</strong> was<br />

then split into divisions depending on the material<br />

their industrial activity was dedicated to (iron,<br />

aluminum and magnesium). In 2002 the aluminum<br />

division was sold to the private equity market to<br />

give birth to <strong>Teksid</strong> <strong>Aluminum</strong>, today’s publisher<br />

of MS&T.<br />

It is known that today’s market dynamics, especially<br />

in the car production industry, is much and<br />

dramatically faster than in the past years, and <strong>Teksid</strong><br />

<strong>Aluminum</strong> had to react to the changing scenario<br />

through a severe restructuring and divestiture<br />

process, which largely resized its traditional<br />

footprint of an international and industry leading<br />

company.<br />

The evolution experienced by <strong>Teksid</strong> <strong>Aluminum</strong><br />

and the radical changes occurred during the last<br />

decades in science, in technology and consequently<br />

in our society, have left the objectives of the journal<br />

unchanged, and its contents have evolved<br />

coherently with the scientific and technologic<br />

trends of metallurgy.<br />

In the Company strategy, these changes have been<br />

accompanied by a constant search for innovative<br />

and competitive product solutions, by a constant<br />

effort to enable the costumers enjoy the company’<br />

s collaboration, from the concept stage to the<br />

delivery of unfinished or completely machined<br />

products and by a constant attention the best<br />

combination of technical performance and<br />

product quality.<br />

This evolution has been reflected by MS&T too.<br />

This is not the place for presenting an exhaustive<br />

list of the topics covered by the journal in all<br />

these years. However probably some readers<br />

will remember academic and technological<br />

contributions dedicated to innovative forming<br />

techniques – such as lost-foam or semisolid<br />

forming, – or to the challenges derived by the<br />

emergence of new structural materials such as cast


magnesium alloys, or to thermomechanical<br />

treatments – like liquid hot isostatic pressing, –<br />

and to new experimental methods for testing metal<br />

properties and performance – thermoelastoplastic<br />

yield stress vs. conventional yield stress, and much<br />

more.<br />

Since its birth, this journal has hosted contributions<br />

of both academic research and metalworking<br />

industry from all over the world. Perhaps the<br />

‘academicians’ have cooperated more<br />

enthusiastically than the industrial researchers to<br />

the “two-way flow of information” originally<br />

wished, but these attitudes is to be attributed both<br />

to the confidential nature of industrial research<br />

and to the old academic caveat “publish or perish”<br />

still alive in the web era..<br />

In all these years <strong>Teksid</strong> <strong>Aluminum</strong> has treasured the MS&T bridging role<br />

between science and industrial expertise in the R&D programs of the Group.<br />

Furthermore the metallurgic community could enjoy a 25-year long support<br />

to scientific divulgation. We can assert that the targets, originally indicated<br />

by Mr. Mosconi, have been reached. Nowadays MS&T papers published since<br />

1983 can be found in university libraries, academic institutions and industrial<br />

R&D laboratories all over the world. Notably, all the issues published since<br />

the year 2000 can be also retrieved comfortably from the Company website<br />

(www.teksidaluminum.com).<br />

The topics such as those mentioned above and unpredictable new ones –<br />

specifically related to light metals and alloys - will continue to arouse interest<br />

of contributors and readers of MS&T for many years to come, as the<br />

increasing cooperation and subscriptions testify.<br />

In a joyful anniversary like this one, we cannot conclude this editorial other<br />

than by congratulating with Metallurgical Science & Technology for its longevity<br />

and scientific maturity, and by wishing Happy Birthday and a long life!<br />

Metallurgical Science and Technology<br />

Editorial Panel


FRACTURE TOUGHNESS AND<br />

MICROSTRUCTURE IN AA 2XXX<br />

ALUMINIUM ALLOYS<br />

1 E. Gariboldi, 1 D. Ripamonti, 1 L. Signorelli, 1 G. Vimercati, 2 F. Casaro<br />

1 Politecnico di Milano, Dipartimento di Meccanica, Milano (MI)<br />

2 Varian Vacuum Technologies, Leinì (TO)<br />

Abstract<br />

The paper presents the toughness properties of forgings made of two AA 2xxx series<br />

aluminium alloys with different microstructural conditions. Fracture toughness tests in<br />

crack opening mode I were performed on compact tension specimens machined from<br />

the forgings in different orientations. The tests were performed both at room temperature<br />

and at 130°C.<br />

Fracture toughness properties were related to microstructural and fractographic features<br />

of the alloys in order to discuss on their failure mechanisms. The effect of the coarse<br />

intermetallic phases within grains or at their boundaries in the different conditions was<br />

underlined. The testing temperature, within the range here investigated, neither affected<br />

fracture toughness properties nor failure mechanisms.<br />

KEYWORDS<br />

Al alloy AA2014, Al alloy AA2618, fracture toughness, microstructure.<br />

3 - Metallurgical Science and Technology<br />

Riassunto<br />

Il lavoro illustra le proprietà di tenacità alla frattura misurate<br />

in forgiati di grandi dimensioni realizzati con due leghe di<br />

alluminio della serie 2xxx in differenti condizioni<br />

microstrutturali. Dai forgiati sono stati ricavati provini CT<br />

con differenti orientazioni sui quali sono state condotte<br />

prove di tenacità a frattura secondo il modo I di apertura<br />

della cricca. Le proprietà ottenute sono state correlate<br />

con la microstruttura riscontrata nei campioni e completate<br />

con analisi frattografiche atte ad individuare i meccanismi<br />

di cedimento. È stato così messo in luce l’effetto delle<br />

diverse microstrutture ed in particolare delle particelle<br />

grossolane di fasi intermetalliche presenti a bordo grano<br />

o all’interno dei grani che differenziano i forgiati nelle leghe<br />

esaminate di composizione più complessa.


1. INTRODUCTION<br />

Aluminium-alloy forging is currently used to<br />

manufacture structural components of relatively<br />

large and complex shape. The plastic deformation<br />

imparted to the material can positively affect its<br />

microstructure by promoting recrystallization<br />

cycles and a greater homogeneity of alloying<br />

elements. However, it should be considered that<br />

in large size forgings, the relatively low amount of<br />

plastic strain given to the alloy cannot completely<br />

refine the structure and intermetallic particles as<br />

in other small-size wrought products such as<br />

extruded bars or rolled sheets. In addition, the<br />

slower quenching rates experienced by large<br />

forgings result in lower mechanical properties<br />

achieved after the subsequent aging process. Large<br />

differences in cooling rate between surface and<br />

centre of large forgings during solution annealing<br />

also result in remarkable residual stresses, that are<br />

often relieved by inserting a plastic deformation<br />

step after quenching and before the aging<br />

treatment [1]. This method also modifies the<br />

precipitation sequence and kinetics of the alloy.<br />

The above described effects significantly affect the<br />

tensile and fracture properties of aluminium alloy<br />

forgings. Focusing the attention on the fracture of<br />

aluminium alloys, it was reported that toughness<br />

is strictly related to the presence of coarse<br />

particles, 0.1 to 10 µm in diameter, that can be<br />

either non-equilibrium particles formed during<br />

2. MATERIALS INVESTIGATED<br />

The present investigation was carried out on three<br />

forgings having a roughly cylindrical shape with a<br />

diameter of 250 mm, made of aluminium<br />

alloys AA2014 (Al4CuSiMg) and AA2618<br />

(Al2Cu1.5MgNi). The parts had been forged from<br />

extruded bars of diameter 190 mm with different<br />

manufacturing cycles.<br />

Two forged samples of the AA2014 alloy were<br />

produced by forging in two steps at 390°C. The<br />

samples were then heat treated to T6 temper by<br />

different parameters. A first sample, hereafter<br />

referred to as 2014-A forging was solution<br />

annealed at 505°C for 6 hours, water quenched<br />

and artificially aged at 160°C for 14 hours, following<br />

the usual industrial heat treatment route. In the<br />

case of the forging in AA2618 alloy (hereafter<br />

referred to as 2618 forging), the solution annealing<br />

at the usual temperature for this material, 530°C,<br />

lasted 1 hour, and it was followed by water<br />

quenching and by artificial aging at 190°C for 20<br />

solidification or inclusions from insoluble impurities [2]. These particles<br />

crack easily as the matrix deforms within the plastic flow zone at the crack<br />

tip and causes the typical ductile fracture mode where crack propagates<br />

via coalescence of voids. The amount, size and distribution of these second<br />

phase particles is thus relevant for the fracture toughness properties and<br />

even material with comparable tensile properties can display significantly<br />

different fracture toughness properties.<br />

In addition to the abovementioned effect, the role played by submicrometer<br />

particles (0.01 to 0.5 µm in diameter) need to be considered. In the case of<br />

the same volume fraction and microstructural features of coarse particles,<br />

a substantial modification of toughness can be observed in age hardenable<br />

aluminium alloys varying the amount and characteristics of fine hardening<br />

particles. The behaviour is complex and the fine hardening particles are<br />

responsible for it. It is well known that the presence of particle having<br />

suitable distribution and size enhances the resistance of peak aged alloys to<br />

deformation and thus tends to reduce the extension of the plastic zone,<br />

positively affecting toughness [2, 3]. On the other hand, the lower strainhardening<br />

capacity of the material in the peak aged condition with respect<br />

to the underaged condition, also gives rise to local plastic instabilities that<br />

significantly contribute to reduce the material toughness [2]. Further, where<br />

grain boundary precipitate free zones are observed, strain localization in<br />

these regions and intergranular ductile fracture can occur [3, 4]. In these<br />

cases the fracture toughness depends on the spacing and size of the voidnucleating<br />

particles at grain boundaries [4, 5]. The higher fracture toughness<br />

displayed by alloys in the underaged with respect to over-aged condition as<br />

well as the transition towards intergranular ductile fracture mode as<br />

overaging proceeds confirm this latter effect [2, 4].<br />

The aim of the present paper is to contribute to a better understanding of<br />

the correlation between microstructure, tensile and toughness properties<br />

of aluminium forgings as a result of different thermomechanical cycles,<br />

focusing in particular to the role of large second phase particles.<br />

hours, following a common industrial practice. The 2014-B forging was<br />

produced and heat treated as for 2618 material, leading to a substantially<br />

overaged matrix and to intermetallic phases distribution rather different<br />

than that of 2014-A forging.<br />

Light optical microscopy observations and Vickers hardness tests (0.98N<br />

load) were performed to evaluate the general microstructural features.<br />

Tensile test specimens were machined from the forgings in the hoop direction<br />

in regions characterized by a homogeneous structure and hardness.<br />

Two sets of Compact Tension (CT) fracture toughness specimens were<br />

machined with cracks laying in diametral planes of the forgings. In the first<br />

set, the crack propagation direction was radial (CR direction according to<br />

ASTM E399-90 and B645-02 [6, 7]) while in the second set it was longitudinal<br />

(CL direction). The specimens had a thickness B of 20 mm [6, 7].<br />

Tests were carried out on a MTS 810 universal testing machine, equipped<br />

with an environmental chamber suitable for test temperatures up to 250°C.<br />

Before fracture toughness tests, a precracking stage was performed at test<br />

temperature until a total crack length (machined notch and fatigue crack)<br />

of <strong>about</strong> 20 mm was reached. Precracking was performed under load control<br />

(sinusoidal cycles at 10 Hz frequency) while crack length was monitored<br />

via the elastic compliance technique by measuring the Crack Opening<br />

Displacement (COD). The stress intensity factor during pre-cracking<br />

decreased linearly with crack length from 11 to 8 MPa•m 1/2 . Fracture<br />

4 - Metallurgical Science and Technology


toughness tests were carried out imposing a displacement rate of 0,025<br />

mm/s, monitoring the COD and applied load until the occurrence of unstable<br />

crack propagation.<br />

In a second stage of the research study, the toughness of the materials at<br />

130°C was investigated. The J IC integral was measured according to the<br />

single-specimen technique and the ASTM E1820-01 standard [8]. The J IC<br />

parameter was evaluated instead of K IC since the material toughness was<br />

expected to be greater than that measured at room temperature. The tests<br />

were performed monitoring crack length via the compliance method at<br />

fixed steps of COD. In order to reduce the strain accumulated under load<br />

during each load step due to creep effects at 130°C, the holding period<br />

under constant applied load (P) for the adjustment of the crack length was<br />

fixed to 1 s. J Q was estimated according to the mentioned standard as the<br />

intercept between the power-law curve (fitting the experimental data within<br />

the range stated by ASTM E1820-01) and the straight line passing to the<br />

3. RESULTS<br />

3.1 MICROSTRUCTURE<br />

All forgings were characterized by grains elongated along the main plastic<br />

flow path experienced during forging. Coarse intermetallic particles were<br />

also present in the microstructure, as expected for these alloys. More<br />

specifically, the 2014-A material (figure 1a) was characterized by elongated<br />

grains with a transversal size of <strong>about</strong> 50 µm and by large intermetallic<br />

particles aligned in the flow direction: globular Al 2 Cu (θ) particles (bright<br />

particles in figure 1a) and blocky shaped clustered particles containing Fe,<br />

Mn, Si and Cu (darker particles in the same figure). In regions where these<br />

particles were observed small equiaxed grains were often detected<br />

A B<br />

C D<br />

5 - Metallurgical Science and Technology<br />

point ∆a = 0.2 mm, P = 0 N with a slope<br />

corresponding to a double flow stress (this latter<br />

corresponding to the average between the yield<br />

and the ultimate tensile stress).<br />

Validity requirements of J 1C tests were not met in<br />

all cases since unstable crack propagation or popin<br />

phenomena occurred in some samples. In these<br />

cases the K Q parameter was evaluated from the<br />

P-COD curves.<br />

Fracture surfaces were observed by scanning<br />

electron microscopy (SEM). Metallographic<br />

sections were also cut perpendicularly to the crack<br />

plane to observe the crack propagation path by<br />

optical microscopy and SEM.<br />

suggesting the local recrystallization effect induced<br />

by these intermetallics.<br />

The direction of plastic flow was less evident in<br />

the 2014-B samples, that displayed rather elongated<br />

grains of 50-100 µm in transverse size together<br />

with smaller equiaxed grains in the regions<br />

containing large intermetallic particles (figure 1b).<br />

These were in lower amount with respect to the<br />

previous alloy. Particles with dark appearance at<br />

grain boundaries (figure 1b) were of complex<br />

chemical composition. On the contrary globular<br />

Al 2 Cu particles had a bright appearance in<br />

Fig. 1:<br />

Typical microstructure of<br />

the investigated forgings.<br />

2014-A (a), 2014-B (b),<br />

2618 at low (c) and<br />

higher (d) magnification.<br />

The longitudinal axis of<br />

forgings is horizontal in<br />

all the micrographs.


micrographs and were observed both inter- and<br />

intragranularly.<br />

The 2618 sample was characterized by very large<br />

grains slightly elongated in the flow direction. Their<br />

size in the transverse direction was of the order<br />

of 500 µm, as shown in figure 1c. As expected,<br />

considering the composition of the AA2618 alloy<br />

[1], several intragranular FeNiAl 9 particles were<br />

observed together with globular Al 2 Cu and (CuFe)<br />

Al 3 precipitates.<br />

The microstructure and hardness of the forgings<br />

were in all cases homogeneous in the regions<br />

where tension and toughness specimens were<br />

sampled.<br />

3.2 MECHANICAL BEHAVIOUR<br />

The average tensile properties in hoop direction<br />

of the investigated materials at room temperature<br />

and at 130°C are summarized in Table I. At room<br />

temperature, the 2014-A forging showed the<br />

highest tensile strength and the lowest ductility.<br />

The increase in test temperature from 20 to 130°C<br />

led to a significant reduction in the ultimate tensile<br />

strength (UTS) and 0,2% proof stress (YS) and to<br />

a corresponding significant increase in reduction<br />

of area (Z) and elongation (A) at fracture.<br />

The results of the fracture toughness tests are<br />

summarized in Table 2 while representative load<br />

vs. COD graphs are shown in Figure 2. Here, the<br />

lines needed for the evaluation of K Q are added<br />

A) B)<br />

6 - Metallurgical Science and Technology<br />

Table 1: Tensile properties of<br />

the investigated alloys<br />

YS [MPa] UTS [MPa] A [%] Z [%]<br />

Temperature 20°C 130°C 20°C 130°C 20°C 130°C 20°C 130°C<br />

2014-A 421 386 454 402 1,8 9,3 3,7 15,8<br />

2014-B 357 341 425 393 7,3 8,4 14,2 16,9<br />

2618 350 345 416 409 6,4 8,6 10,8 12,3<br />

Table 2: Fracture toughness of the materials<br />

investigated (K Q or K IC according to ASTM B645<br />

and K JIC accordiing to ASTM E1820)<br />

Specimen Temperature 2014-A 2014-B 2618<br />

direction [°C] MPa . m Ω MPa . m Ω MPa . m Ω<br />

CL 20 19,3 (K IC ) 23,0 (K Q ) 22,4 (K Q )<br />

CL 20 20,1 (K IC ) 24,9 (K Q ) 23,3 (K Q )<br />

Average CL 20 19,7 (K IC ) 24,0 (K Q ) 22,9 (K Q )<br />

CL 130 18,8 (K IC ) 23,4 (K JIC ) 21,9 (K Q )<br />

CR 20 23,4 (K IC ) 24,8 (K Q ) 26,9 (K Q )<br />

CR 20 22,8 (K IC ) 24,4 (K Q ) 26,7 (K Q )<br />

Average CR 20 23,1 (K IC ) 24,6 (K Q ) 26,8 (K Q )<br />

CR 130 21,4 (K JIC ) 23,4 (K JIC ) 25,6 (K Q )<br />

to the experimental curves<br />

The condition of plain strain crack propagation in CT specimen fracture<br />

can be met for sufficiently large specimens. The minimum specimen thickness<br />

Fig. 2: Load vs. COD plots of the samples tested at 20°C.<br />

A) 2014-A forging, CL specimen; B) 2014-B forging, CR specimen


Table 3: F values for the materials and specimen<br />

orientation investigated<br />

Specimen Temperature 2014-A 2014-B 2618<br />

direction [°C]<br />

CL 20 9.4 4.8 4.9<br />

CL 20 8.7 4.1 4.5<br />

Average CL 20 9.1 4.5 4.7<br />

CL 130 9.1 4.2 4.9<br />

CR 20 6.4 4.1 3.4<br />

CR 20 6.8 4.3 3.4<br />

Average CR 20 6.6 4.2 3.4<br />

CR 130 7.4 3.3 3.6<br />

Table 4: J IC values (N/mm) of the investigated<br />

forgings tested at 130°C<br />

Specimen Temperature 2014-A 2014-B 2618<br />

direction [°C] [N/mm] [N/mm] [N/mm]<br />

CL 130 - 7.23 -<br />

CR 130 5.96 7.00 -<br />

(B) to be adopted can be calculated, at the end of the tests, according to<br />

the two standard above considered:<br />

B min =5·(K Q /YS) 2 according to ASTM B645 (1)<br />

B min =2,5·(K Q /YS) 2 according to ASTM E399 (2)<br />

A) B)<br />

Fig. 2: Load vs. COD plots of the samples tested at 20°C.<br />

A) 2014-A forging, CL specimen; B) 2014-B forging, CR specimen<br />

7 - Metallurgical Science and Technology<br />

Thus, the minimum thickness required by the E399<br />

standard is half of that required by the B645<br />

standard for aluminium alloys. The agreement to<br />

plain strain conditions and the possible deviation<br />

from the minimum value of B can be evaluated<br />

using a parameter F, defined as:<br />

F=B·(YS/K Q ) 2 (3)<br />

Where F greater than 5 means that the<br />

requirements for plain strain condition are met<br />

for both standards and valid K IC can be obtained.<br />

2,5


according to the ASTM E1820 standard. As far as<br />

the AA2618 alloy that resulted to be the toughest<br />

material at room temperature, at 130°C it showed<br />

unstable crack propagation (Figure 3b) that<br />

prevented the evaluation of J IC. In the cases of valid<br />

J IC , K JIC was computed, assuming plain strain<br />

conditions, by using the following correlation<br />

proposed by ASTM E1820 standard:<br />

3.3. FRACTOGRAPHIC<br />

OBSERVATIONS<br />

Fractographic observations were performed at<br />

two main locations on fracture surface of each<br />

specimen in the crack propagation region: the first<br />

next to, the latter at 4 mm from the boundary<br />

with fatigue precrack. Selected images at low<br />

magnification of the first location for the different<br />

alloys and specimen orientation are reported in<br />

figure 4. In the unstable crack propagation region,<br />

the 2014-A forging showed the typical features of<br />

ductile fracture, generated by nucleation of dimples<br />

mainly from the coarser intermetallic particles<br />

fractured in a brittle way (figure 5a). Within the<br />

matrix, regions with dimples of far smaller size<br />

were also visible, laying on well-defined planes<br />

A)<br />

C)<br />

K JIC = [(E/(1-v 2 ))•J IC ] 0.5 (4)<br />

K JIC values are also listed in Table 2 for comparison purpose. Examination of<br />

this table clearly shows that toughness of 2014-B forging is greater than<br />

that of the same alloy in forging 2014-A. Further, it can be stated that in all<br />

the examined samples, crack propagation in CL orientation is favoured<br />

with respect to that in CR orientation.<br />

(figure 5b). These can be correlated to the presence of fine particles that in<br />

some cases were observed inside the dimples in high magnification images<br />

that suggest ductile intergranular fracture.<br />

The fracture surfaces of the 2014-B forging did not show significant<br />

differences from 2014-A material, with relatively extended regions of<br />

microdimples (figure 7a) that could be correlated to the presence of coarse<br />

grain boundary particles (figure 6b). The size and distribution of these<br />

particles corresponded to those observed at grain boundaries (figure 4c).<br />

The 2618 forging, showed a transgranular ductile fracture mode (figures<br />

4d, 7b). Dimples observed on fracture surfaces nucleated from the<br />

homogeneously distributed FeNiAl 9 particles (figures 6c and 6d).<br />

The fracture surfaces of specimens tested at 130°C of forgings made of<br />

AA2014 alloy were similar to those tested at room temperatures (figure<br />

8a). On the contrary, as the temperature increased, small-size microdimples<br />

were observed on the fracture surface of the forging made of AA2618 alloy<br />

(compare figure 7b and 8b).<br />

Fig. 4: Fracture surface of cracks propagated at room temperature in the boundary region between fatigue precracking and unstable crack<br />

propagation region. 2014-A forging, CL (a) and CR (b) directions. 2014-B forging; CR direction (c) and 2618, CR direction (d).<br />

B)<br />

D)<br />

8 - Metallurgical Science and Technology


A) B)<br />

A)<br />

C)<br />

Fig. 5: Fracture surface of cracks propagating at room temperature in 2014-A forging in CL (a) and CR (b) direction.<br />

Fig. 6: Microstructure in the region of unstable crack propagation on longitudinal section of CT specimens. a) 2014-A forging, CL direction;<br />

b) 2014-B forging, CL direction; c) and d) 2618 forging, CL and CR directions, respectively.<br />

B)<br />

D)<br />

A) B)<br />

Fig. 7: Fracture surface sampled in CL direction of 2014-B forging (a) and of 2618 forging (b) tested at room temperature.<br />

9 - Metallurgical Science and Technology


A) B)<br />

Fig. 8: Fracture surface of cracks propagated at 130°C in specimen with CL orientation of 2014-A (a) and 2618 (b) forgings.<br />

4. DISCUSSION<br />

In the examined forgings crack propagated in a<br />

ductile manner, in some cases by an intergranular<br />

ductile mode, linking the early fractured coarse<br />

intermetallic particles, located at inter or<br />

intragranular position in the different materials<br />

investigated.<br />

Following the approach proposed by Hahn and<br />

Rosenfield [2] for aluminium alloys, it can be stated<br />

that crack propagates when the size of extensive<br />

plastic strain formed ahead of the crack tip<br />

corresponds to the average coarse interparticle<br />

distance (δ) Further, according to these authors,<br />

K and d are correlated as follows:<br />

δ =(0.5K 2 )/(E•YS) (5)<br />

In the case of 2618 forging the δ values estimated<br />

using room temperature tensile and toughness<br />

properties were 11 and 15 µm for specimen<br />

orientations CL and CR, respectively, comparable<br />

to the average interparticle size between the<br />

intragranular particles (mainly FeNiAl 9 particles)<br />

in longitudinal and radial direction, respectively. The<br />

preferred particle orientation and their tendency<br />

to align axially in the forging regions where<br />

specimens were sampled corresponds to a greater<br />

CONCLUSIONS<br />

The toughness properties in opening mode I of<br />

forgings made of AA2014 and AA2618 aluminium<br />

alloys with different microstrctural conditions were<br />

presented.<br />

Fracture toughness ranged between 19 and 26<br />

MPam 0.5 , depending on the material and specimen<br />

direction. Fracture toughness was in general lower<br />

interparticle distance in the crack propagation direction of CR specimens.<br />

As previously presented, in the specimens obtained from forgings made of<br />

AA2014 alloy the crack proceeds in a ductile manner but along an<br />

intergranular path, intercepting the oriented coarse aligned intermetallic<br />

particles axially. The values of δ computed from equation 5 for these two<br />

forgings are 6.7 and 9 µm for CL and CR specimen of 2014-A forging,<br />

respectively. In the case of the last forging (2014-B), δ equals 12 µm for<br />

both CL and CR directions. Thus, also in the case of AA2014 alloy, d is<br />

comparable to the interparticle distance along the crack path. It can be<br />

thus stated that fracture toughness is correlated to the mean interparticle<br />

distance along the crack path.<br />

The above observations well agree with the simple correlation proposed<br />

by Hahn and Rosenfield [2] for the case of forged Al-alloy parts (where<br />

fracture is initiated by cracking of large intermetallic particles once reached<br />

critical stress/strain levels in the region of extensive deformation at the<br />

crack tip [3]) and fracture toughness is correlated to the mean interparticle<br />

distance along the crack path.<br />

Thus, forgings of the same heat treatable alloy, even in the same heat<br />

treatment condition and with comparable hardness and tensile properties<br />

can show significantly different fracture behaviour, depending on intermetallic<br />

particle population. Especially, as the mean interparticle distance along the<br />

crack path between the coarser intermetallic particles increases, the<br />

nucleation of voids at these particles is shifted at higher applied loads and<br />

toughness is improved.<br />

The role of coarse intermetallic particles and the need to reduce their<br />

amount, to optimize their size and distribution in forgings (taking into account<br />

the most critical crack propagation directions in the components) in order<br />

to increase material toughness is therefore highlighted.<br />

for specimen sampled in CL direction. In the examined forgings crack<br />

propagated in ductile manner, in some cases by an intergranular ductile<br />

mode, linking in any case the voids formed at the early fractured inter- or<br />

intragranular coarse intermetallic particles. No significant difference in<br />

toughness nor in the fracture mode was observed when tests were<br />

performed at room temperature or at 130°C.<br />

Fracture toughness properties were related to the presence and distribution<br />

of the coarse intermetallic phases within grains or at their boundaries in<br />

the different investigated alloys.<br />

10 - Metallurgical Science and Technology


The simple correlation proposed by Hahn and Rosenfield between fracture<br />

toughness and the mean interparticle size was found to be applicable when<br />

the interparticle distance along the crack path (different for different sampling<br />

direction in forgings where these particle were aligned along flow directions)<br />

is taken into account. An observation and quantitative assessment of the<br />

REFERENCES<br />

[1] J.S. Robinson, R.L. Cudd, J.T. Evans. Creep resistant aluminium alloys<br />

and their applications. Material Science and Technology, 19 (2003) pp.<br />

143-155.<br />

[2] G.T. Hahn, A.R. Rosenfield. Metallurgical Factors Affecting Fracture<br />

Toughness of <strong>Aluminum</strong> Alloys. Metall. Trans. A, (1975) pp. 653-668.<br />

[3] T. Kobayashi. Strength and fracture of aluminium alloys. Materials Science<br />

Engineering, A286 (2000) pp. 333-341.<br />

[4] A.P. Reynolds, E.P. Crooks. The effect of thermal exposure on the fracture<br />

behaviour of aluminium alloys intended for elevated temperature<br />

service. in “Elevated temperature on fatigue and fracture”, ASTM STP<br />

1297. Williamsbourg, Virginia(USA), (1997) 191-205.<br />

[5] B.Q. Li, A.P. Reynolds. Correlation of grain-boundaries precipitates<br />

parameters with fracture youghness in an Al-Cu-Mg-Ag alloy subjected<br />

to long-term thermal exposure. Journal of Materials Science, 33 (1998)<br />

pp. 5849-5853.<br />

[6] ASTM B 645-02 Practice for plane-strain fracture toughness testing of<br />

aluminium alloys.<br />

[7] ASTM E 399-99 Standard test method for plane-strain fracture<br />

toughness of metallic materials.<br />

[8] ASTM E1820-01. Standard test method for measurement of fracture<br />

toughness.<br />

11 - Metallurgical Science and Technology<br />

distribution of coarse intermetallic phases on<br />

optical micrographs can be a useful tool to check<br />

fracture toughness properties of different forgings<br />

in their peak aged condition when yield strength<br />

is known.


VARIATIONS IN MICROSTRUCTURE AND<br />

MECHANICAL PROPERTIES OF CAST<br />

ALUMINIUM EN AC 43100 ALLOY<br />

Salem Seifeddine, Torsten Sjögren and Ingvar L Svensson<br />

Jönköping University, School of Engineering, Component technology - Sweden<br />

Abstract<br />

The microstructure and mechanical properties of a gravity<br />

die and sand cast Al-10%Si-0.4%Mg alloy, which is one of<br />

the most important and frequently used industrial casting<br />

alloys, were examined. Tensile test samples were prepared<br />

from fan blades and sectioned through three positions<br />

which experienced different cooling rates. Furthermore,<br />

the inherent strength potential of the alloy was revealed<br />

by producing homogeneous and well fed specimens with a<br />

variety of microstructural coarseness, low content of oxide<br />

films and micro-porosity defects, solidified in a laboratory<br />

environment by gradient solidification technology. The<br />

solidification behaviour of the alloy was characterized by<br />

thermal analysis. By means of cooling curves, the<br />

solidification time and evolution of the microstructure was<br />

recorded. The relation between the microstructure and<br />

the mechanical properties was also assessed by using quality<br />

index-strength charts developed for the alloy. This study<br />

shows that the microstructural features, especially the ironrich<br />

needles denoted as β-Al FeSi, and mechanical<br />

5<br />

properties are markedly affected by the different processing<br />

routes. The solidification rate exerts a significant effect on<br />

the coarseness of the microstructure and the intermetallic<br />

compounds that evolve during solidification, and this<br />

directly influences the tensile properties.<br />

KEYWORDS<br />

Aluminium cast alloys, gravity- and sand casting,<br />

gradient solidification technique, thermal analysis,<br />

porosity, quality index.<br />

Riassunto<br />

In questo lavoro sono state esaminate la microstruttura e le proprietà meccaniche di<br />

una delle leghe più comuni nell’industria delle leghe leggere, la lega Al-10%Si-0.4%Mg,<br />

colata per gravità sia in conchiglia che in sabbia. Le provette di trazione sono state<br />

ricavate da pala di ventilatore e sono state sezionate in tre posizioni caratterizzate da<br />

diverse velocità di raffreddamento. Inoltre è stato valutato il valore potenziale di resistenza<br />

a trazione ottenibile della lega, mediante produzione di campioni caratterizzati da diversa<br />

morfologia strutturale, basso contenuto di film ossidi microporosità, solidificati in<br />

ambiente controllato con metodologia di solidificazione a gradiente. Il comportamento<br />

in solidificazione è stato caratterizzato mediante analisi termica. Per mezzo di curve di<br />

raffreddamento, si è determinato il tempo di solidificazione e l’evoluzione della<br />

microstruttura. La relazione tra la microstruttura e le proprietà meccaniche è stata<br />

determinata usando specifiche carte di correlazione indice di qualità – resistenza. Questo<br />

studio dimostra che le caratteristiche microstrutturali, in particolare i grani aciculari ad<br />

alto tenore di ferro indicati come β-Al5FeSi, e le proprietà meccaniche sono<br />

marcatamente condizionate dai differenti percorsi preparativi. La velocità di<br />

raffreddamento esercita una significativa influenza sulla morfologia microstrutturale e<br />

sui componenti intermetallici che si sviluppano durante la solidificazione, influenzando<br />

così direttamente le proprietà di resistenza a trazione.<br />

12 - Metallurgical Science and Technology


INTRODUCTION<br />

The mechanical properties of cast aluminium alloys are very sensitive to<br />

composition, metallurgy and heat treatment, the casting process and the<br />

formation of defects during mould filling and solidification. The coarseness<br />

of the microstructure and the type of phases that evolve during solidification<br />

are fundamental in affecting the mechanical behaviour of the material. The<br />

mechanical properties obtained from gradient solidified samples are assumed<br />

to represent the upper performance limits of a particular alloy, due to the<br />

low level of defects obtained from the favourable melting and solidification<br />

procedure.<br />

Different processing routes offer different qualities and soundness of<br />

commercial components. The true potential of most alloys is seldom<br />

approached, and there exists a lack of data in the literature describing the<br />

variations in microstructure and related mechanical properties of alloy EN<br />

AC 43100. The present investigation was therefore carried out in order to<br />

Fig. 1: Fan blade cast in sand- and gravity moulds.<br />

13 - Metallurgical Science and Technology<br />

elucidate and assess the influence of the casting<br />

process and resultant variations of microstructure<br />

on the mechanical properties, by extracting tensile<br />

specimens from Al-10%Si-0.4%Mg commercial cast<br />

components manufactured by two different casting<br />

methods; sand and gravity die casting. In addition,<br />

the inherent strength potential of this casting alloy<br />

will be investigated by manufacturing tensile test<br />

specimens using gradient solidification technology.<br />

The mechanical properties which have been<br />

measured in this work are: ultimate tensile strength,<br />

yield strength and fracture elongation. The<br />

microstructural features that have been evaluated<br />

are the secondary dendrite arm spacing (SDAS),<br />

pore fraction and precipitated phases such as the<br />

length of the iron-rich β-phase, Al 5 FeSi.<br />

MATERIALS AND<br />

EXPERIMENTS<br />

Cast component<br />

The components that have been studied are fan<br />

blades, as shown in figure 1, which are produced<br />

commercially in a great variety of shapes and sizes,<br />

and by different casting methods. This particular<br />

component is considered to be of interest to study<br />

and analyse due to the widespread use of different<br />

casting methods to manufacture fan blades, which<br />

in this case were manufactured by sand mouldand<br />

gravity die casting.<br />

The locations of the specimens extracted for study<br />

are shown in figure 1; from the top (T), middle<br />

(M) and bottom (B) parts of the fan blade. These<br />

locations were chosen due to their different<br />

solidification times (i.e. thicknesses), which results<br />

in different coarseness of the microstructure,<br />

intermetallic phases and defects.<br />

Cast alloy<br />

Table 1: Chemical composition of the casting alloy<br />

Composition Elements (wt. %)<br />

The commercial blades and specimens produced<br />

by gradient solidification experiments were cast<br />

using alloy EN AC 43100. The standard<br />

composition of the alloy and the actual<br />

composition are given in table 1. Strontium was<br />

Alloy Standard Si Fe Cu Mn Mg Zn Ti Cr Ni Al<br />

EN AC-43100 SS-EN 1706 9.0-11.0 0.55 0.10 0.45 0.20-0.45 0.10 0.15 - 0.05 Bal.<br />

EN AC-43100 Actual sample 9.96 0.47 0.11 0.25 0.35 0.097 0.018 0.01 0.01 Bal.


added in the form of Al-10%Sr master alloy at a<br />

level of 150- 200 ppm.<br />

Sand and gravity die casting<br />

The sand moulds used were made by hand. As a<br />

pattern, a fan blade cast in a gravity mould was<br />

used, which is as already mentioned, one of the<br />

methods normally used to manufacture this size<br />

and type of fan blade. The sand mould consisted<br />

of chemical bonded sand.<br />

Gradient solidification<br />

The same alloy as that used for the sand and gravity<br />

die cast components, shown in table 1, was cast<br />

into rods for further solidification studies<br />

in the gradient solidification equipment. The<br />

gradient solidification technique allows the<br />

production of a high quality material with a low<br />

content of oxide films, porosity and shrinkage<br />

related defects, and produces a homogenous and<br />

A)<br />

well-fed microstructure throughout the entire specimen. In this work a<br />

resistance-heated furnace with an electrically driven elevator was used, see<br />

figure 2a. Three different growth velocities, v, were used, 0.03 mm/s, 0.3<br />

mm/s and 3 mm/s, which correspond respectively to an SDAS of <strong>about</strong> 50,<br />

20 and 7 µm. At each velocity, three specimens were made. The specimens<br />

cast in the gradient solidification equipment were protected by argon in<br />

order to avoid oxidation.<br />

Thermal analysis<br />

In order to study the thermal history of the solidified components,<br />

thermocouples were inserted at three different positions in the sand cast<br />

fan blade, as indicated in the picture in figure 1. To record the cooling curves<br />

a data acquisition device from National Instruments was used. In each of<br />

the sand moulds three thermocouples were placed. The thermocouples<br />

were of S-type, which means that the different materials in the thermocouple<br />

wires are platinum and platinum-rhodium (90Pt-10Rh). When manufacturing<br />

the thermocouples, the wires are inserted in a two-hole Al 2 O 3 -tube to<br />

separate the wires, which are connected to a plug. The thermocouples<br />

were protected from the molten metal by a quartz glass tube. The maximum<br />

operating temperature for the S-type thermocouple is over 1500°C [1].<br />

The recording rig is illustrated in figure 2b.<br />

B)<br />

Fig. 2: . Illustration of the instruments that have been used during the investigation. While a) demonstrates the gradient solidification equipment,<br />

b) presents the data acquisition rig that is connected to the sand moulds.<br />

14 - Metallurgical Science and Technology


RESULTS OF THE<br />

MICROSTRUCTURE AND<br />

MECHANICAL PROPERTIES<br />

INVESTIGATION<br />

Since specimens were extracted from fan blades<br />

with different thickness and at different sections<br />

it was necessary to machine them in order to<br />

obtain specimens suitable for tensile testing. The<br />

gradient solidified specimens have also been<br />

prepared in the same manner, and the dimensions<br />

of the samples used for tensile testing are shown<br />

in figure 3. In order to analyse the microstructure<br />

and measure the different microstructural features,<br />

a light microscope, image processing software and<br />

a scanning electron microscope, SEM, were used.<br />

Microstructure development<br />

The microstructure coarseness is defined as the secondary dendrite arm<br />

spacing, SDAS, which is also a function of the local solidification time, and<br />

quantifies the scale of the microstructure and its constituents. Examining<br />

the microstructure of specimens from both sand and gravity die casting, it<br />

has been observed that there is a wide scattering of SDAS and consequently<br />

variations of sizes and shapes of the intermetallic compounds within the<br />

same component. The needle-shaped β-phase, Al 5 FeSi also seems to coexist<br />

with the α-phase, Al 15 (Fe,Mn) 3 Si 2 , which has the morphology of Chinese<br />

script. As depicted in figure 4, the dendrites are equiaxed and the eutectic<br />

is finely modified. Comparing the microstructure formations, it can be noticed<br />

that in the case of gravity die casting the microstructure is finer, due to the<br />

higher cooling rate. Furthermore, the different coarsenesses of<br />

microstructure, the phases and their morphologies are comparable to the<br />

sand cast microstructure.<br />

It is relevant to note that even if Mn has been added to the alloy to act as<br />

A) B)<br />

Fig. 3: A schematic illustration of the tensile specimen (dimensions in mm).<br />

Fig. 4: . a) Illustration of the microstructure of a gravity die cast specimen while b) shows a sand cast microstructure. The black spot is a pore.<br />

<br />

<br />

15 - Metallurgical Science and Technology<br />

<br />

<br />

<br />

an Fe-corrector and to promote the formation of<br />

Chinese script, see table 1, iron-rich Al 5 FeSi-needles<br />

are still formed. These needles act as local stress<br />

raisers in the matrix and are considered to be<br />

deleterious for the mechanical performance, which<br />

is why their formation and growth should be taken<br />

into consideration and as much as possible<br />

suppressed.<br />

In order to understand the local variations in<br />

microstructure, and to understand the formation<br />

sequences and development of constituent phases<br />

in the sand cast component, thermal analysis was<br />

implemented. When manufacturing the sand cast<br />

fan blades at the foundry, data for the cooling<br />

curves was recorded. In each of the sand moulds,<br />

three thermocouples were used, as shown in figure 1.


With this composition, see table 1, and casting<br />

technique, three different cooling curves have been<br />

obtained and are presented in figure 5a. The figure<br />

also shows that there exists clear variations in local<br />

solidification times which are directly related to<br />

the variations in component thickness, see also<br />

table 2. These variations in cooling time lead to<br />

major local differences in microstructures which<br />

affect the tensile performance of the cast<br />

components. The SDAS for the sand cast and<br />

gravity die cast component varied between 47-61<br />

µm and 20-45 µm respectively.<br />

Figure 5b shows that the freezing range is<br />

approximately between 595°C and 545°C. From<br />

the different cooling curves it is possible to see at<br />

which temperature and time the dendritic network<br />

starts to grow, at which temperature and time the<br />

eutectic starts to grow and also the total<br />

solidification time, from pouring the melt until the<br />

fraction solid has reached 100%.<br />

When the value of the derivative exceeds zero,<br />

figure 5b, the temperature is actually increasing<br />

due to the release of latent heat due to phase<br />

transformations. This is most evident in the case<br />

where a positive derivative is observed at <strong>about</strong><br />

<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Mechanical properties<br />

The results of the tensile tests from the<br />

commercial fan blades and the laboratory<br />

prepared samples are shown in the diagrams in<br />

590°C when the dendritic network starts to grow. The next positive value<br />

of the derivative appears when the eutectic starts to grow, at <strong>about</strong> 570°C.<br />

When the temperature is around 550°C precipitations of a complex eutectic<br />

with intermetallics and a hardening phase such as Mg 2 Si starts to precipitate<br />

from the remaining melt. From the different thermocouples the following<br />

solidification times have been determined:<br />

Table 2. Solidification times for the different parts<br />

of the sand cast fan blade.<br />

Thermocouple Thinnest part Mid part Bottom<br />

(T) (M) part (B)<br />

Solidification time (s) 260 470 610<br />

Material thickness (mm) 12 20 28<br />

Examining the gradient solidified specimens, a significant discrepancy between<br />

the different solidification rates is observed. The higher the cooling rate<br />

the finer the microstructure, as illustrated in figure 6. The iron-rich<br />

β-phase Al 5 FeSi is found to be very short and very well dispersed for the<br />

two highest solidification velocities; 3 and 0.3 mm/s, with a corresponding<br />

SDAS ≈ 7 and 23 µm respectively, in comparison to the lowest velocity<br />

0.03 mm/s, with SDAS ≈ 47 µm, where long iron-bearing phases were found<br />

randomly distributed in the eutectic regions.<br />

Fig. 5: The thermal history of the sand cast component is registered in figure a) while in b) an analysis of the cooling curve<br />

obtained from the bottom section is performed.<br />

figure 7. The two different casting methods are compared in figure 7a, where<br />

each point represents a different position on the fan blade. A wide scatter<br />

in data is clearly observed but it can be noticed that the properties of the<br />

thinner parts seem to approach higher combinations of stress and strain<br />

values.<br />

16 - Metallurgical Science and Technology


In order to simulate the processes of the cast components, tensile test<br />

samples of the same alloy and corresponding microstructural coarseness<br />

have been produced by using gradient solidification. By producing a more<br />

homogenous microstructure with a lower level of defects, the material’s<br />

inherent potential was revealed, see figure 7b. The scatter in data in figure<br />

7b is ascribed to different levels of defects, and the higher stress-strain<br />

A) B)<br />

<br />

Fig. 6: Phases and microstructures for a) SDAS = 47 µm obtained at 0.03 mm/s and b) SDAS = 7 µm obtained at 3 mm/s.<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 7: The graph in a) illustrates the ultimate tensile strength and elongation to fracture for the gravity die and sand casting specimens where B is<br />

the bottom (thickest) section, M the middle and T the top (thinnest) part of the blade. The graph in b) presents the tensile properties for the<br />

gradient solidified specimens with a range of cooling rates.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

17 - Metallurgical Science and Technology<br />

<br />

combinations are the limits for what this alloy<br />

might perform under ideal conditions.<br />

When quantifying the SDAS, five measurements<br />

were taken at three randomly chosen positions of<br />

the specimen. From these data a mean value was<br />

calculated. At high solidification rates the SDAS<br />

µ<br />

µ<br />

µ


appears to be very small and the microstructure<br />

has been found to be more uniform. As noticed,<br />

the overall effect of obtaining a small SDAS is an<br />

improvement in both ductility and tensile strength,<br />

see figure 8.<br />

When measuring the length of the iron-rich Al 5 FeSi<br />

needles (β-phase), the 15 longest needles observed<br />

in the microstructure were measured, considering<br />

the entire cross-section area of the sample, and a<br />

mean value was assessed. It should be mentioned<br />

that the needles were not chemically analyzed, but<br />

<br />

<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

<br />

µ<br />

<br />

<br />

were identified according to their morphologies and colours. A correlation<br />

was found between the length of the β-phase and the mechanical properties.<br />

An increase in needle length results in a reduction of tensile strength and<br />

elongation as can be seen in figure 9. Note that the iron-bearing β-phase<br />

was too short to be measured or not present at all in the specimen solidified<br />

at a rate of 3 mm/s, SDAS = 7 µm. It is also seen that even if these needles<br />

are relatively shorter in some cases than in corresponding samples, they<br />

exhibited lower properties. The premature fractures must therefore be<br />

due to other defects such as oxide film inclusions or other brittle and<br />

undesirable phases such as Al 8 Mg 3 FeSi 6 . These intermetallics are harmful to<br />

the mechanical properties since cracking of these phases reduces the ductility,<br />

<br />

Fig. 8: The graphs a) and b) illustrate the influence of SDAS on the tensile properties.<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ<br />

<br />

<br />

<br />

<br />

µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

18 - Metallurgical Science and Technology<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

<br />

<br />

µ<br />

Fig. 9: The influence of the β-phase needle length on a) ultimate tensile strength and b) the elongation to fracture.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ


µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

Fig. 10: This graph shows the relation between the β-phase length and the SDAS.<br />

and their formation reduces the Mg available in the melt to form Mg 2 Si<br />

which has a hardening effect and improves both the yield and ultimate<br />

tensile strengths.<br />

Generally, larger SDAS is associated with lower cooling rates and also with<br />

longer β-phase needles. The local solidification time influences the growth<br />

of the intermetallics that develop and grow during the solidification process,<br />

and hence a relation between SDAS and the β-phase length is observed as<br />

shown in figure 10.<br />

When a tensile sample contains pores it is reasonable to assume that the<br />

region of porosity will yield first due to the reduced load bearing area<br />

concentrating the stress near the voids. The area fraction of porosity was<br />

determined by measuring the area just below and at the fracture surface.<br />

<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 11: The graphs a) and b) illustrate the influence of porosity on the tensile properties.<br />

<br />

19 - Metallurgical Science and Technology<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

The first mentioned was measured by an image<br />

analyzer and the latter was photographed and<br />

observed by scanning electron microscopy, SEM.<br />

This method enables an examination of the<br />

distribution and size of the pores.<br />

In the present study, it has been observed that the<br />

cooling rate, SDAS in this case, seem to govern<br />

the length as well as the percentage area fraction<br />

of porosity to some extent, see figure 11. It appears<br />

that a reduction in SDAS results in smaller average<br />

pore size and a reduced area fraction of porosity.<br />

It is obvious that as the solidification time is low,<br />

less time will therefore be available for the diffusion<br />

of the hydrogen into the interdendritic regions<br />

which results in small sizes and fraction of pores.<br />

But comparing the porosity formation, the gradient<br />

solidified samples with SDAS ~ 23 µm are<br />

associated with largest percentage area fraction<br />

of porosity which might be due to the arrestment<br />

of premature pores as they become entrapped by<br />

the advancing solidification front.<br />

The discrepancy could be derived from the theory<br />

that involves the harmfulness of the oxide films,<br />

which suggests that when an oxide particle is<br />

approached by the solidification front, the particle<br />

experiences the hydrogen-rich environment<br />

produced by the rejection of gas from the<br />

advancing solid. Furthermore, the access of gas by<br />

diffusion into the air pocket in the gap of the oxide<br />

particles, the pore will start to form and grow. As<br />

the freezing rate is slow and the particle is poorly<br />

wetted by the melt, time will therefore be available<br />

for more hydrogen to diffuse resulting in pore<br />

expansion and growth.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ


In the case of the larger SDAS ~ 47 µm, it is<br />

therefore reasonable to assume that due the slow<br />

cooling conditions many air pockets formed within<br />

the liquid or due to interaction with oxide particles<br />

have been driven in front of the solidified front<br />

without being engulfed and in that case not been<br />

detected in the gradient solidified specimens due<br />

to the solidification mode. Another reason might<br />

be the longer time available for the hydrogen to<br />

diffuse and move in front of solid-liquid interface<br />

out of the sample into the surrounding<br />

environment.<br />

The scatter in the data is too great to derive any<br />

clear correlation between porosity level and the<br />

tensile strength and ultimate elongation, see figure<br />

11 a and b. According to literature, it is stated [2]<br />

that the static tensile properties such as ultimate<br />

tensile strength, yield strength and elongation to<br />

A) B)<br />

DISCUSSION<br />

Generally, the mechanical properties of metals are<br />

widely influenced by their microstructures. When<br />

defining the microstructure many parameters have<br />

to be taken into account, and these include among<br />

others the secondary dendrite arm spacing, the<br />

size and shape of precipitated phases such as silicon<br />

and iron-bearing phases, grain size, porosity, etc.<br />

In this investigation, many of the parameters<br />

mentioned above have been measured and<br />

correlated to the mechanical properties.<br />

Constituents that may have deleterious effect on<br />

the mechanical properties are defects such as<br />

oxide films or undesired phases and inclusions.<br />

A comparison of the results of the tensile testing<br />

with the casting method and the corresponding<br />

fracture, were all decreased with an increased degree of porosity. On the<br />

contrary, according to [3-7], the reduction in tensile properties has almost<br />

no correlation with the average volume fraction of porosity. In fact, the<br />

decrease in tensile properties was attributed to the length and/or the area<br />

fraction of defects in the fracture surface. Development of a high fraction<br />

of porosity may also be due to Fe in the melt. The long, needle-shaped iron<br />

intermetallic phase formed is expected to cause severe feeding difficulties<br />

during solidification. The morphology of the β-phase blocks the interdendritic<br />

flow channels, which is why it is proposed that higher iron contents in the<br />

alloy are associated with higher levels of porosity.<br />

SEM examination was also performed in order to study what kinds of pores<br />

are frequently found on the fracture surfaces in this work. Figure 12a<br />

illustrates a pore seen in a sand cast specimen and 12b is seen in a gradient<br />

solidified specimen at a solidification rate of 3 mm/s.<br />

Worth to indicate is that the melt hydrogen content has not been measured<br />

and assumed in this case to be constant for all the alloys since they have<br />

been produced under similar conditions.<br />

Fig. 12: Illustration of pores. a) Showing sphericity of a possible gas pore b) The cavity is observed at the edge of the<br />

fractured surface shows a possible gas-shrinkage pore.<br />

specimens with three different solidification rates obtained by gradient<br />

solidification shows clearly that the tensile behaviour and production method<br />

is related by three different stress-strain curves. The material exhibiting<br />

SDAS ≈ 7 µm results in the highest stress-strain combinations as shown in<br />

the diagram in figure 13. The main reason for this behaviour is the finer<br />

microstructure, a more homogeneous distribution of the intermetallic phases<br />

and lower levels of porosity. Even though there are no specimens with<br />

corresponding SDAS extracted from commercial components available in<br />

this study, this data clearly shows the inherent potential strength of the<br />

alloy.<br />

The stress-strain values for the sand cast specimens follow the curve for<br />

SDAS = 47 µm (obtained by gradient solidification), but fracture occurs at<br />

lower stress and strain levels, see figure 13. The stress-strain curve for<br />

SDAS = 23 µm (by gradient solidification) corresponds well to the stressstrain<br />

values of the gravity die cast specimens.<br />

The reason why the sand and gravity die cast specimens fractured much<br />

earlier than the gradient solidified specimens, and exhibited different<br />

20 - Metallurgical Science and Technology


µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 13: This graph shows tensile test curves for the gradient solidified specimens and<br />

includes mechanical properties data for the gravity die and sand cast specimens.<br />

combinations of strength and fracture elongation values may be due to the<br />

formation of brittle and coarse intermetallics such as the iron-bearing<br />

β-phase and an inhomogeneous microstructure. Even if similar SDAS has<br />

been obtained from the gradient solidification equipment, the variations in<br />

the β-phase needles, local coarsenesses of microstructures and sizes of<br />

intermetallics are likely to be different due to microsegregation in the<br />

castings. Segregation of elements is likely to occur in castings and this will<br />

affect the development and solidification sequences of the microstructure<br />

and lead to local variations in mechanical properties. At the thinner part of<br />

the fan blades the β-phase needles are likely to be short and have no time<br />

to grow and coarsen as is the case in thicker parts where the local iron<br />

<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

21 - Metallurgical Science and Technology<br />

level might also be different. It has also been shown<br />

that a finer microstructure with smaller SDAS is<br />

beneficial for the mechanical properties. A finer<br />

microstructure means larger areas of grain<br />

boundary leading to more difficulties for<br />

continuation of slip and dislocation movement<br />

during deformation, hence yielding stronger alloys.<br />

A simple tool that might be used to predict the<br />

impact of changes in the chemical composition,<br />

solidification rates and routes, microstructure and<br />

heat treatment on the tensile performance is by<br />

implementing a quality index, Q. Its practical value<br />

originates from the fact that the mechanical<br />

“quality” of an alloy is expressed by a single<br />

numerical parameter, whose value can be read off<br />

a quality index chart or calculated through a simple<br />

equation relating the elongation to fracture, s , the f<br />

ultimate tensile strength, UTS, and d, which is an<br />

empirical parameter chosen to be around 150 MPa<br />

for EN AC 43100 (Al-7%Si-0.4%Mg) to make Q<br />

more or less independent of the yield strength of<br />

the alloy, as seen in equations 1-2, [8-10]:<br />

Q = UTS + d * log (100*s ) (1)<br />

f<br />

n<br />

σ = Kε<br />

(2)<br />

where σ is the true flow stress, K is the alloy’s<br />

strength coefficient, ε is the true plastic strain and<br />

n the strain hardening exponent.<br />

In order to produce the charts in figure 14, a mean<br />

value for the strength coefficient, K, has been<br />

calculated. For the commercial cast components,<br />

figure 14a, K is approximately 410 MPa and for<br />

the gradient solidified materials K is around 430<br />

MPa, figure 14b. The strain hardening exponent, n,<br />

is numerically equal to the necking onset strain<br />

Fig. 14: A quality index chart for the alloy studied obtained for gravity- and sand cast samples is presented in a) while b) represents the gradient<br />

solidified specimens. The mean value of the strength coefficient is a little higher for the gradient solidified specimens.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

µ<br />

µ<br />

µ


and can be used to define the relative ductility<br />

parameter, q, see equation 3.<br />

q = s f /n (3)<br />

Defining q = 1 means the onset of necking<br />

represents samples that fail either at the onset of<br />

necking or beyond, while q < 1 identifies lower<br />

quality or less ductile samples. For instance q =<br />

0.5, 0.2, 0.1 identifies samples that fail at 50%, 20%<br />

and 10% of their maximum uniform strain,<br />

respectively.<br />

As observed, the points corresponding to<br />

SDAS = 52 µm are located near the bottom of<br />

the chart, which may be due to their coarser<br />

CONCLUSIONS<br />

As an outcome from the current investigation, the<br />

dendritic cell size, SDAS, and the length of ironbearing<br />

phase whether as needles or Chinese<br />

script, governs the tensile strength and ductility of<br />

Al-Si-Mg alloys. The early fracture of sand and<br />

gravity die cast specimen can be related to the<br />

length of the iron-bearing phase, oxide films<br />

entrapment, and inhomogeneous microstructure<br />

and to the Si particle size and morphology, which<br />

determine the rate of particle cracking with the<br />

applied strain.<br />

• A finer microstructure, small and well<br />

distributed iron-rich compounds and low levels<br />

of porosity exert a beneficial impact on the<br />

resulting mechanical properties.<br />

• The locally increased levels of iron due to<br />

ACKNOWLEDGEMENTS<br />

REFERENCES<br />

1. Metals Handbook; ASM International, 1998, pp.<br />

663-665.<br />

2. M. Avalle, G. Belingardi, M.P. Cavatorta and R.<br />

Doglione, International Journal of Fatigue Vol. 24<br />

, (2002), pp. 1-9.<br />

3. C.H. Cáceres and B.I. Selling , Materials Science<br />

and Engineering A, Vol. 220, Issues 1-2, (1996),<br />

pp. 109-116.<br />

4. M. Harada, T. Suzuki and I. Fukui , Journal of the<br />

Japan foundrymen´s society Vol. 55, No. 12,<br />

(1983), pp. 47-50.<br />

microstructure and the length of the β-phase. As the strength of the gravity<br />

die cast specimens increases, the data points shift toward the upper right<br />

corner.<br />

Regarding the gradient solidified samples, the data points for the<br />

SDAS = 7 µm specimens, with the finer microstructure, exhibit the highest<br />

values of Q and q. These values seem to be the only ones that reach the<br />

limit imposed by the necking line q = 1.<br />

As the SDAS and the level of deleterious microstructural compounds are<br />

decreased, the strength and ductility of the alloy is increased and so does<br />

the Q, allowing for a possible comparison between different alloys and<br />

processing routes. Furthermore, the Si particles, playing a role as reinforcing<br />

phases, decrease in size and the shape becomes more fibrous, which has an<br />

overall beneficial effect on the mechanical properties.<br />

possible segregations might have resulted in longer and larger fractions<br />

of β-phase needles. The main reason for the remarkable reduction of<br />

the mechanical performance of commercial components is proposed<br />

to be due to the presence of the locally induced stresses by the needles<br />

which break, link and propagate the cracks in an unstable manner. The<br />

higher the solidification rate the shorter the length of the needles and<br />

the sounder the material.<br />

• The gradient solidification method shows that the EN AC 43100 alloy<br />

used for commercial cast components processed by sand and gravity<br />

die casting, has an improvement potential to provide sounder castings<br />

with substantial ductility.<br />

• The addition of Mn did not alter the morphology of iron-bearing βphase<br />

into a more compact Chinese script. Instead, these iron<br />

intermetallics seem to be coexisting in the microstructure.<br />

• Porosity is not found to be the primary parameter controlling the tensile<br />

strength and ductility, but it is worth bearing in mind that the reduction<br />

of the load bearing area generally has a harmful influence on the overall<br />

properties.<br />

The authors acknowledge Johan and Jan Tegnemo at Mönsterås Metall AB for their invaluable help and support for making the<br />

experiments in the foundry and Nordic Industrial Fund for the financial support within the Nordic Project STALPK.<br />

5. M.K. Surappa, E. Blank and J.C. Jaquet, Scripta Metallurgica, Vol. 20, No. 9,<br />

(1986), pp. 1281-1286.<br />

6. J.G. Conley, J. Huang, J. Asada and K. Akiba, One Hundred Third Annual<br />

Meeting of the American Foundrymen’s Society; Rosemont, IL; USA, (1998).<br />

pp. 737-742.<br />

7. C.H . Cáceres, Scripta Materialia Vol. 32 No.11, (1995), pp. 1851-1856.<br />

8. C.H. Cáceres, I.L. Svensson, J.A. Taylor, International Journal of Cast Metals<br />

Research, Vol. 15, No. 5, (2003), pp. 531-543.<br />

9. C.H. Cáceres, International Journal of Cast Metals Research, Vol.12, (2000),<br />

pp. 367-375.<br />

10. M. Drouzy, S. Jacob and M.Richard, AFS International Cast Metals Journal,<br />

Vol. 5, No. 2, (1980), pp. 43-50.<br />

22 - Metallurgical Science and Technology


COMPARATIVE STUDY OF HIGH<br />

TEMPERATURE WORKABILITY OF ZM21<br />

AND AZ31 MAGNESIUM ALLOYS<br />

1 M. El Mehtedi, 1 L. Balloni, 1 S. Spigarelli, 1 E. Evangelista, 2 G. Rosen, 3 B.H.Lee, C.S.Lee<br />

1 Dipartimento di Meccanica, Università Politecnica delle Marche, Ancona, Italy<br />

2 ALUBIN, Kiryat Bialik, Israel<br />

3 Department of Materials Science and Engineering, Pohang University, Pohang, Korea<br />

Abstract<br />

High temperature regime, 300-450°C for Mg-Al-Zn alloys, is currently used in primary<br />

processing, such as rolling and extrusion, as well as for secondary operation like forging.<br />

The knowledge of temperature and strain rate proper combination (processing window)<br />

as well as the microstructure evolution occurring during hot deformation clarifies the<br />

relationships between forming variables and final properties of components. Numerous<br />

data on AZ31 and few other Mg-Al alloys, produced by laboratory testing, are available<br />

in the scientific and technical literature. The ZM21, Mg-2Zn-1Mn, by contrast, is<br />

characterized by absolute lack of scientific data. In the alloy the addition of manganese,<br />

by suppressing the formation of beta phase, increases the solidus temperature that<br />

results in the larger processing window than in AZ31. The benefit requires extensive<br />

analysis aimed at optimizing the deformation variables that affect the microstructure<br />

refinement under dynamic and static recrystallization. The high-temperature plastic<br />

deformation and the microstructure evolution of the ZM21 were thus investigated in<br />

the temperature range between 200 and 500°C and results were analysed and compared<br />

with those of a conventional heat-treated AZ31.<br />

23 - Metallurgical Science and Technology<br />

Riassunto<br />

Il regime delle alte temperature 300-450°C, è comunemente<br />

usato per le operazioni primarie di deformazione plastica,<br />

come la laminazione e l’estrusione, come pure per quelle<br />

secondarie come la forgiatura o lo stampaggio delle leghe<br />

Mg-Al-Zn. La conoscenza della combinazione adatta<br />

temperatura-velocità di deformazione (finestra di processo)<br />

e della evoluzione della microstruttura fa luce sulle relazioni<br />

tra variabili di deformazione e proprietà dei componenti.<br />

Numerosi dati, ottenuti con prove di compressione o<br />

torsione, sono disponibili sulla lega AZ31 e su pochi altre<br />

leghe della classe Mg-Al. Sulla lega ZM21 (Mg-2Zn-1Mn)<br />

non sono invece disponibili dati scientifici. Nella lega<br />

l’addizione del Mn sopprime la formazione della fase beta<br />

ed aumenta la temperatura del solidus, producendo una<br />

finestra di processo più ampia di quella dell’AZ31. Questi<br />

aspetti richiedono analisi approfondite sono rivolte<br />

all’ottimizzazione delle variabili di deformazione che poi<br />

governano l’affinamento della microstruttura per effetto<br />

della ricristallizzazione sia dinamica che statica. La<br />

deformazione plastica ad alta temperatura e l’evoluzione<br />

della microstruttura della lega ZM21 sono state analizzate<br />

nell’intervallo di temperatura 200-500°C e i risultati sono<br />

confrontati con quelli di una lega convenzionale trattata<br />

termicamente.


1. INTRODUCTION<br />

Lightweight constructions are becoming more and<br />

more important for the automotive industries due<br />

to increase of fuel prices and legislative<br />

requirements like CAFÉ in the US, and the directive<br />

or the control of CO 2 emissions in the EU [1,2].<br />

Magnesium alloys for their unique combination of<br />

light weight, high specific strength and stiffness and<br />

high recycling capability are candidate to replace<br />

steel, iron and, in some cases, also aluminum parts.<br />

The growing interest in magnesium alloys is<br />

expected to increase in the next years [3]. The<br />

use of components manufactured via die-, sand-,<br />

mould- and rheo-casting is fairly well introduced<br />

in a wide number of applications which are ranging<br />

from steering wheels, instrumental panels, gear box<br />

housing and even to hybrid engine blocks. The<br />

scenario for the automotive industry is that the<br />

use of magnesium alloys for many components,<br />

mainly as castings, is further expanding in the short<br />

term, whereas the application in body and chassis<br />

components, including sheet and extrusions, is<br />

anticipated in the medium term [4].<br />

A driving force for the introduction of wrought<br />

alloys is their property enhancement, since casting<br />

alloys are affected by porosity, which would be of<br />

particular importance for structural and crashrelevant<br />

parts. Initial applications for this category<br />

are expected for extrusions and forgings, to be<br />

followed by sheets. These requirements may<br />

address researches towards the modification of<br />

known alloys as well as the development of new<br />

ones aiming at improving both mechanical<br />

performances, also by ageing (T6), and high<br />

temperature stability. The growing interest in use<br />

of magnesium alloys is expected to increase in the<br />

next years, and to be sustained by research [5].<br />

From recent magnesium alloys conferences [6], the<br />

following items stand out as limiting or retarding<br />

factors for the use of wrought alloys: 1) poor cold<br />

forming properties; 2) limited number of wrought<br />

alloys and lack of processing data. As far as the<br />

item 1) is concerned, it is important to note that<br />

magnesium alloys are difficult to work at low<br />

temperature (


EXPERIMENTAL<br />

The AZ31 experimental result data belong to the authors’ wide database<br />

obtained in the last 3 years by testing a batch of AZ31 with different initial<br />

microstructure: ingot, extruded, heat treated, overaged. These investigations<br />

aimed at clarifying the microstructure features that significantly affect the<br />

high temperature deformation of the alloy; the data were partly published<br />

[11,12] and used to identify the processing window (strain rate and<br />

temperature conditions) that is successfully used for industrial forming<br />

operations. Warm forming experiments were also carried out to assess the<br />

AZ31 sheet formability [13]. The present data were obtained by testing<br />

the specimens machined from rolled AZ31 heat treated at 385°C for 14 h,<br />

at the Pohang University.<br />

The ZM21 alloy was produced and extruded by Alubin, Israel. The specimens<br />

for torsion testing, with gauge radius R, and length L, 5 and 10 mm respectively,<br />

were machined from extruded rods. Torsion tests were carried out in air<br />

on a computer-controlled torsion machine, under equivalent strain rates<br />

ranging from 10 -3 to 5 s -1 and temperatures from 200°C to 400°C. Specimens<br />

were heated 1°C/s by induction coils and maintained 5 minutes before<br />

torsion to stabilize the testing temperature that was measured by<br />

thermocouple in contact with the gauge section. Specimens just after the<br />

fracture were rapidly quenched with water jets to avoid microstructure<br />

RESULTS AND DISCUSSION<br />

The initial microstructure of the alloys before testing is displayed in Figure<br />

1. The heat-treated AZ31 exhibited an equiaxed and fine microstructure.<br />

The microstructure of the extruded ZM21 is composed by duplex grain<br />

size, the small and the large grains were 20 and 60 µm respectively.<br />

Figures 2 and 3 show typical equivalent stress vs. equivalent strain curves<br />

obtained by testing at 200 and 400°C the rolled and heat treated AZ31 and<br />

the extruded ZM21, respectively. The flow curves shape of the rolled AZ31<br />

does not differ appreciably from that observed in the cast [11,12]. At<br />

A) B)<br />

25 - Metallurgical Science and Technology<br />

modifications during slow cooling.<br />

The von Mises equivalent stress, σ, and equivalent<br />

strain, ε, were calculated using the relationships:<br />

σ =<br />

3 M<br />

( 3 + m'+<br />

n'<br />

)<br />

2 π R<br />

3<br />

2 π N R<br />

ε =<br />

2)<br />

3 L<br />

where N is the number of revolutions, M is the<br />

torque, m’ (strain rate sensitivity coefficient) at<br />

constant strain is ∂ log M / ∂ log N,<br />

•<br />

and n’(strain<br />

hardening coefficient) at constant strain rate is<br />

∂ log M / ∂ log N; at the peak stress, clearly n’=0.<br />

To reveal the microstructure, longitudinal sections<br />

at the periphery of the sample gauge, i.e. in the<br />

point where the strain and strain rates assumed<br />

the values calculated by equations 1) and 2), were<br />

investigated by means of optical microscopy.<br />

200°C, under all strain rates, the flow curves of<br />

both alloys exhibit well defined σ y , ~60 MPa for<br />

AZ31 and ~20 MPa for ZM21, a continuous<br />

increase of work hardening up to σ max , which is<br />

larger in AZ31, followed by fracture that in ZM21<br />

occurs at larger strain. At 400°C, AZ31 the flow<br />

curve, after yielding and very limited workhardening,<br />

gets a steady state region, typical<br />

behaviour of dynamic recovery (DRV), up to<br />

fracture. The ZM21 flow curves display yielding<br />

Fig. 1: Microstructure of the rolled and heat-treated AZ31 (a) and of the extruded ZM21 (b).<br />

1)


and rapid hardening up to σ peaks located at ε~0.7,<br />

irrespective of the different strain rates, followed<br />

by a continuous softening towards the fracture<br />

that occurs at e larger than in AZ31; at 400°C and<br />

strain rate 0.05 s -1 the maximum strain value is<br />

2.5. The flow curve exhibiting a peak followed by<br />

softening in stationary state is indicative of dynamic<br />

recrystallization (DRX) that nucleates in proximity<br />

of the peak producing new stable grains whose<br />

size controls the softening and the σ of stationary<br />

state.<br />

In the present instance, the absence of stationary<br />

state and the decrease of σ up to fracture may be<br />

indicative of three processes promoting the<br />

continuous softening: (i) the onset of DRX around<br />

the peak, producing new small grains, (ii) the fast<br />

new-grain growth at the critical size for (iii) crack<br />

nucleation and growth.<br />

A significant decrease in the strain to failure in<br />

torsion can be observed (Figure 4), in particular<br />

in the high-temperature region; an additional<br />

difference is the observation of a well defined<br />

yielding in ZM21, (Figure 3) i.e. of the presence of<br />

a short strain range without load increase, an effect<br />

A)<br />

<br />

B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

A)<br />

B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 3: Equivalent stress vs. equivalent strain curves obtained in torsion for<br />

ZM21 at 200 (a) and 400°C (b).<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

26 - Metallurgical Science and Technology<br />

Fig. 2: Equivalent stress vs. equivalent strain curves obtained in<br />

torsion for AZ31 at 200 (a) and 400°C (b).<br />

probably due to a sudden increase in the number of<br />

dislocations unpinned from their solute atmosphere<br />

(either Zn or Mn).<br />

The temperature and strain rate dependence of the<br />

peak stress are shown in Figure 5. The experimental<br />

data are well described by the equation:<br />

ε& = A[ sinh(<br />

ασ)<br />

] n<br />

exp( −Q<br />

/ RT)<br />

3)<br />

where A and α are material parameters, n is a<br />

constant, R is the gas constant, T is the absolute<br />

temperature, and Q is the apparent activation energy<br />

for high-temperature deformation process. In the<br />

AZ31 a best-fitting procedure gave α=0.017 MPa-1 and n= 4.4 for temperatures between 200°C and<br />

400°C. The activation energy, calculated by plotting<br />

sinh(as) as a function of 1/T, was Q=140 kJ/mol. The<br />

value is relatively close to the one of self-diffusion of<br />

Mg (135 kJ/mol) or for diffusion of Al in Mg (143 kJ/<br />

mol). In the case of the ZM21, α was close to 0.049<br />

MPa-1 , and n ranged between 2.6 and 4.4 (average<br />

n=3.3). The activation energy for high temperature<br />

deformation was close to 200 kJ/mol, higher than in<br />

the case of AZ31.


A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 4: Equivalent strain to failure for AZ31 (a) and ZM21 (b).<br />

A) B)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ασ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 6: Zener-Hollomon parameter as a function of peak-flow stress for AZ31 and<br />

ZM21.<br />

27 - Metallurgical Science and Technology<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Fig. 5: Peak flow stress dependence on applied stress as a function of temperature for AZ31 (a) and ZM21 (b).<br />

<br />

<br />

<br />

<br />

ασ<br />

α <br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

ασ<br />

Figure 6 plots the Zener-Hollomon parameter, Z,<br />

defined as<br />

[ ( ) ] n<br />

sinh<br />

<br />

<br />

<br />

<br />

<br />

Z = ε&<br />

exp( Q / RT)<br />

= ασ<br />

4)<br />

as a function of peak stress, with Q= 170 kJ/mol<br />

and α=0.034 MPa-1 (mean values of those obtained<br />

experimentally for AZ31 and ZM21). Analysis of<br />

Figure 6 clearly demonstrates that the peak flow<br />

stresses are higher in the AZ31 than in the ZM21.<br />

Figure 7 shows typical microstructure of the AZ31<br />

quenched after torsion testing. At 200°C, and a<br />

strain rate of 10-2 s-1 , the structure is mostly<br />

composed by slightly elongated grains (Figure 7a),<br />

with very fine recrystallized grains on several grain<br />

boundaries. At 300°C, under the same strain rate,<br />

the recrystallized fraction is substantially higher<br />

than at the lower temperature, and the average<br />

size of the recrystallized grains is larger (Figure


A) B)<br />

C) D)<br />

Fig. 7: Microstructure of the AZ31 tested at: 200°C-10-2 s-1 (a) 300°C-10-2 s-1 (b) and 400°C- 1 s-1 (c); (d) 400°C-10-3 s-1 .<br />

7b). Fully recrystallization and localised grain<br />

growth at 400°C, in the high strain rate range,<br />

resulted in the presence of some large grains<br />

heavily deformed by twinning, albeit a relatively<br />

large fraction of the structure is composed by fine<br />

equiaxed grains. Under lower strain rate, the<br />

structure is composed by a homogeneous<br />

distribution of relatively coarse equiaxed grains<br />

(Figure 7d).<br />

The mechanism of dynamic recrystallization (DRX)<br />

has been analysed by McQueen and Konopleva<br />

[15]; twinning in fact takes place before the other<br />

mechanisms, except basal glide that occurs in<br />

favourably oriented grains. At low strains, twinning<br />

favourably reorients grains for slip. As deformation<br />

reaches a sufficiently high value, DRX starts where<br />

high misorientations have been created by<br />

accumulation of dislocations, i.e. where slip has<br />

occurred on several slip system (near grain<br />

boundaries and twins). The new fine grains form a<br />

mantle (or a “necklace”) along grain boundaries<br />

and deform more easily than the grain core, thus repeatedly undergoing<br />

recrystallization [12].<br />

Figure 8 shows typical micrographs of the microstructure of ZM21 tested<br />

in torsion. The structure, even at 300°C, is largely unrecrystallized, with<br />

elongated grains; at 400°C, the grains are mostly equiaxed, but elongated<br />

structures still persist. The strong decrease in flow stress after the peak<br />

should be attributed to the late onset (at strain close to ε=0.7-0.8) of<br />

recrystallization and coarsening phenomena, that in this alloy exhibit quite<br />

different kinetics than in the case of the AZ31. Only at the highest<br />

temperature (500°C), the microstructure underwent complete<br />

recrystallization and grain growth, a process that resulted in the presence<br />

of coarse (50 µm) grains.<br />

The above analysis gives some interesting preliminary indications on the<br />

high-temperature formability of the extruded ZM21 alloy. In particular:<br />

1. the intrinsic ductility, i.e. the equivalent failure strain in torsion, is lower<br />

in ZM21 than in AZ31;<br />

2. the fraction of recrystallized structure, at all the investigated<br />

temperatures, is lower in ZM21 than in AZ31;<br />

3. the maximum equivalent stress, i.e. the working load, is higher in AZ31<br />

than in ZM21; yielding has been observed in ZM21, thus suggesting a<br />

solute strengthening role played by either Zn or Mn;<br />

28 - Metallurgical Science and Technology


A)<br />

C)<br />

Fig. 8: Microstructure of the ZM21 tested at: 300°C-5 s-1 (a) 400°C- 5x10-2 s-1 (b) and 500°C- 5 s-1 (c).<br />

4. the different composition of the ZM21 enables the use of higher<br />

working temperature, up to 500°C, i.e. well above the upper limit<br />

allowed for AZ31. On the other hand, even though strain to failure is<br />

relatively high at 500°C, the microstructure undergoes extensive grain<br />

CONCLUSIONS<br />

The high temperature workability of the extruded ZM21 alloy has been<br />

investigated by torsion tests between 200 and 500°C. Both mechanical and<br />

microstructural results were compared with those observed in a rolled<br />

AZ31 alloy; the comparison indicated that the strain to fracture in torsion<br />

were significantly lower in ZM21 than in AZ31, even though in general in<br />

the former alloy the ductility exceeded the minimum allowable value of<br />

ε=1. These differences in mechanical response were attributed to a lower<br />

tendency to dynamic recrystallization; while in rolled AZ31 early dynamic<br />

recrystallization was observed at temperatures as low as 200°C,<br />

microstructural observations clearly suggested that in the extruded ZM21<br />

the lower limit for DRX was shifted toward significantly higher temperatures.<br />

B)<br />

29 - Metallurgical Science and Technology<br />

growth, that precludes the fine grain size that<br />

has been observed to be a prerequisite for<br />

optimum sheet formability.<br />

ACKNOWLEDGEMENTS<br />

The authors greatly acknowledge the Ministries<br />

of Foreign Affairs of Italy and Korea who in the<br />

frame of “2004-06 Joint Scientific and Technological<br />

Research Program” promoted the research by<br />

funding the researchers’ mobility.<br />

The valid support of D. Ciccarelli in experimental<br />

activity has been appreciated.


REFERENCES<br />

1) CAFÉ, http:/www.nhtsa.dot.gov<br />

2) Communication from the Commission to the<br />

Council and European Parliament:<br />

Implementing the Community strategy to<br />

reduce CO emissions from cars. Third annual<br />

2<br />

report (reporting year2001), COM (2002)<br />

693.<br />

3) EU Directive on the End of Life of Vehicles,<br />

2000/53/EC.<br />

4) S. Schumann, Mat. Sci. Forum 488-489 (2005),<br />

1-8.<br />

5) Ya Zhang, X. Zeng, C. Lu, W. Ding, Y. Zhu, Mat.<br />

Sci. Forum, 488-489,(2005), 123.<br />

6) Magnesium Technology 2006, A.A. Luo et al,<br />

eds. TMS, 2006.<br />

7) F.J. Humphreys, M. Hatherly: Recrystallization<br />

and related annealing phenomena. Pergamon<br />

Press, Oxford, 1996.<br />

8) J. Koike, Mat. Sci.Forum, 419-422 (2003), 199.<br />

9) H. Utsunomiya, T. Sakai, S. Minamiguchi, H. Koh, Magnesium Technology<br />

2006, A.A. Luo et al. eds. TMS, 2006, 201.<br />

10) J. Bohlen, J. Swoistek, W.H. Sillekens, P.J. Vet, D. Letzig, K.U. Kainer,<br />

Magnesium Technology 2005, TMS, 2005, 241.<br />

11) S. Spigarelli, M. El Mehtedi, E. Evangelista, J. Kaneko, Metall.Sci.Techn. 23<br />

(2005) 11-17.<br />

12) S. Spigarelli, M. El Mehtedi, M. Cabibbo, E. Evangelista, J. Kaneko, A. Jäger,<br />

V. Gartnerova, mater. Sci.Eng. A 462 (2007), 197.<br />

13) A. Forcellese, M. El Mehtedi, M. Simoncini, S. Spigarelli, Key Engineering<br />

Materials 344 (2007) 31.<br />

14) S. Spigarelli, E. Evangelista, M. El Mehtedi, L. Balloni, Mechanical And<br />

Microstructural Aspects Of High Temperature Formability Of AZ31<br />

Sheets, Proceeding 10 ESAFORM 2007, Zaragoza, Spain, 1305.<br />

15) H.J. McQueen, E.V. Konopleva, in Magnesium Technology 2001, J. Hryn<br />

ed., TMS, 2001, p. 227.<br />

30 - Metallurgical Science and Technology


INSTRUCTIONS FOR AUTHORS<br />

To ensure that space can be provided for the publication of several papers in<br />

each issue of the Metallurgical Science and Technology, Authors are kindly<br />

requested, to limit their contributions to a maximum of fifteen typed pages<br />

prepared in accordance with the following instructions:<br />

1. Texts must be submitted on Word for Windows compatible CD in the<br />

following format: double line space, and divided into paragraphs; each page<br />

should contain 30-35 lines, each composed of 50-60 characters and spaces.<br />

Authors are also asked to supply a reserve copy on A4 paper in the same<br />

format to cover the possibility that their disc may not be legible.<br />

2. Each paragraph should be indented three spaces.<br />

3. Notes must be numbered and typed on separate sheets.<br />

4. All measurements must be expressed in ISO 1000-81 units (International<br />

System).<br />

5. Please set out the following details clearly at the end of the typescript: first<br />

name and surname, university department or firm where employed, full<br />

address and telephone number. This information will help the editorial staff<br />

to get in touch with you if necessary.<br />

6. A short summary should be provided (not more than 150 words). This<br />

should give an account of the scope of the paper and summarise its<br />

conclusions.<br />

7. Tables, graphs, and figures must be set out on separate sheets. Their position<br />

should be indicated in the text in the form “table 1”, “fig. 3”, etc. If data<br />

collected by other persons are used in their preparation, acknowledgement<br />

of the source should be made in a note at the foot of the table or figure<br />

concerned. The titles of tables, graphs, figures, etc. must provide a specific<br />

indication of their contents.<br />

All photographs should be supplied in their original form. Tables, graphs and<br />

figures are recommended in JPEG format (300 dpi) and in grey scale.<br />

8. Bibliographical references should be set out on a separate sheet or sheets<br />

in the order in which they appear in the text. The following rules must be<br />

followed:<br />

a) Authors’ names: surname of the first author, comma, initials, comma,<br />

initials and surnames of remaining authors; “and” between the final pair<br />

of authors; finished with full point.<br />

b) Article title: initial caps and lower case (except in German, where nouns<br />

are always capitalized); finished with full point.<br />

c) Journal name: underlined in the typescript; abbreviations may be used;<br />

initial caps; comma; issue number underlined followed by year in<br />

parentheses, comma; page number (without “p”), final full point.<br />

For book titles the style is similar as follows:<br />

a) Authors’ names: as for journals.<br />

b) Chapter title: as for article title.<br />

c) Book title: “In” followed by initials, surname of editor and “(Ed.)”, comma<br />

and then title underlined, comma.<br />

d) Publisher, comma, location, comma, year, comma, volume, part, etc., comma.<br />

e) Page number preceded by “p”, or “pp.” and followed by full point.<br />

Examples:<br />

Slatcher, S., and J.F. Knott. The ductile fracture of high strength steels. In D.<br />

Francois (Ed.), Advances in fracture mechanics, Pergamon Press, Oxford, 1981,<br />

vol. I, pp. 201-207.<br />

Griffiths, J.R. and D.R.J. Owen. An elastic-plastic analysis for a notched bar in<br />

plane strain bending. J. Mech. Phys. Solids, 19 (1971), 419-431.<br />

The authors are strongly recommended to submit their papers on a CD<br />

prepared using the Word for Windows programs, besides the written<br />

version.


ISSN 0393-6074

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!