You are on page 1of 10

Acta Materialia 51 (2003) 3233–3242

www.actamat-journals.com

Deformation behaviour of iron-rich iron-aluminium alloys at


high temperatures
J. Herrmann 1, G. Inden, G. Sauthoff ∗
Max-Planck-Institut für Eisenforschung GmbH, Max-Planck-Str.1, 40237 Düsseldorf, Germany

Received 8 January 2003; received in revised form 21 February 2003; accepted 10 March 2003

Abstract

The deformation behaviour of binary monocrystalline and polycrystalline Fe-Al alloys with Al contents up to 18
at.% and only low unavoidable impurity contents—in particular less than 100 wt.ppm C—has been studied above room
temperature up to 800 °C. The effects of quenching and annealing treatments on the behaviour of as-cast materials
were investigated in order to clarify the dependence of strength and ductility on Al content and possible short-range
ordering. Besides solid-solution hardening by Al, serrated yielding and deformation twinning were observed at high
temperatures. Yielding is affected by quenched-in excess vacancies.
 2003 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.

Keywords: Metal & Alloys (iron alloys); Ordering (short range); Mechanical properties (brittle-to-ductile transition, Portevin-Le Chat-
elier effect, twinning)

1. Introduction and ductility [12–14]. Strength and ductility are


further affected by vacancies [15–19], which are
Fe-Al alloys are of high interest because of readily produced by quenching and which show a
reduced density and high hot-corrosion resistance low mobility [20–24]. The atomic ordering pro-
[1–3] and have been chosen as a base for duces the intermetallic iron-aluminides Fe3Al and
developing new steels for automotive applications FeAl which are used for developing new intermet-
[4,5]. Fe-Al alloys have been studied for a long allic alloys for structural high-temperature appli-
time and it is known that the alloying with Al pro- cations [9,18,25,26].
duces hardening [6,7] and reduces ductility [2,6– Alloys with Al contents in the range of 10–18
9]. At higher Al contents atomic ordering occurs at.% Al are of particular interest since complex
affecting the elastic behaviour [10,11] and strength short-range ordering may occur leading to the so-
called K-state [27]. Recent work has concentrated
on the deformation behaviour of such binary Fe-

Corresponding author. Tel.: +49-211-6792-313; fax: +49- Al alloys at room temperature as a function of Al
211-6792-537.
E-mail address: sauthoff@mpie.de (G. Sauthoff). content and prior heat treatment [28]. It was found
1
Now at Sulzer Innotec, PB Box 65, 8404 Winterthur, that the deformation behaviour of such Fe-Al
Switzerland alloys is controlled primarily by solid-solution

1359-6454/03/$30.00  2003 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
doi:10.1016/S1359-6454(03)00144-7
3234 J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242

hardening with negligible effects of possible short-


range ordering. The present work extends this
work to higher temperatures above room tempera-
ture. The effects of ternary alloying additions,
which are transition metals or C and Si, are
reported separately [29,30]. Preliminary results
have been presented previously [31]. Further
details are available in ref. [32].

2. Experimental

Binary Fe-Al alloys and monocrystals with Al


contents in the range 4–18% Al (always at.%) were
prepared, heat-treated, analysed and studied by Fig. 1. Compressive stress-strain curves (10⫺4 s⫺1 com-
optical light-microscopy and transmission electron pression rate) at various temperatures for polycrystalline as-cast
Fe-17Al with 20 ppm C (—) and 50 ppm C (- - -).
microscopy (TEM) as described before [28]. In
addition to the cast alloys with coarse-grained by the accidentally microalloyed carbon of 20 and
microstructures, a hot-rolled Fe-Al alloy with 15% 50 ppm, respectively. Obviously, the higher C level
Al and 39 ppm C (always wt.ppm) from Thyssen leads to higher yield stresses and hardening rates.
Krupp Stahl AG was used, which was recrystal- This different behaviour is most pronounced at 500
lised at 900 °C to establish a fine-grained micro- °C. In addition the alloy with the higher C content
structure. shows serrated yielding at 300 and 400 °C for plas-
The alloys were subjected to compressive stress- tic strains above 5%. The other alloy with only 20
strain testing as described before [28]. Toughness ppm C shows much finer serrations at higher
was studied by Charpy impact testing using speci- strains and furthermore serrations of quite different
mens of dimensions 3 × 4 × 27 mm3 with notch character at 500 and 590 °C.
according to German standard DIN EN 10045. The Fig. 2 summarises the data for 4 alloys with dif-
used impact tester allowed for recording of the
impact force (as given by the hammer strain) and
specimen flexure (as given by the integrated ham-
mer velocity) as a function of time [33]. The
brittle-to-ductile transition temperature BDTT was
determined as mean temperature in the transition
range between low and high impact energy values.

3. Results

3.1. Strength and ductility

The stress-strain behaviour at higher tempera-


tures above room temperature is illustrated by Fig.
1 for as-cast Fe-17%Al. Clearly both the yield Fig. 2. Compressive yield stress (쎲) and upper yield stress
stress and the strain-hardening rate decrease with (䊏) at 10⫺4 s⫺1 compression rate as a function of temperature
for various polycrystalline as-cast Fe-Al alloys with 50–60 ppm
increasing temperature as is expected. In addition, C with concurrent twinning at low [28] and intermediate tem-
a yield stress drop is visible at 590 °C. The two peratures and serrated yielding at intermediate and higher tem-
alloys in Fig. 1 differ by the impurity content, i.e. peratures.
J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242 3235

ferent Al contents and C contents of 50–60 ppm. stress, and it is even lower than for the as-cast state
Deformation twinning, which was observed at in the case of the Fe-17%Al alloy in Fig. 3. Yield-
lower temperatures below 0 °C for the higher Al ing occurs with coarse serrations and additional
content of 14% [28,32], was found at 200 °C for deformation twinning was found occasionally as
Fe-17%Al. The range of serrated yielding at inter- revealed by metallographic observations (Fig.
mediate temperatures with a yield stress plateau for 4(a)). This high-temperature deformation twinning
Fe-8%Al and reduced yield stress decrease for was mostly observed in furnace-cooled specimens
higher Al contents is centred on the temperature of (Fig. 4(a)) whereas quenched specimens (Fig. 4(b))
300 °C. At 600 °C and higher temperatures yield showed deformation twinning only rarely.
stress drops were observed for the higher Al con-
tents. 3.2. Impact toughness
The specifics of yield stress, serrated yielding
and deformation twinning in the intermediate tem- The toughness of the polycrystalline alloys was
perature range do not depend only on Al content, studied as a function of temperature by Charpy
but also on prior heat treatment and possibly on impact tests which allowed the recording of impact
the presence of carbon as is exemplified by Fig. hammer force and displacement as a function of
3. Obviously the quench from 1100 °C leads to time. Fig. 5 shows typical force-displacement
increased yield stresses in the intermediate tem- curves which reveal the characteristics of the
perature range of 200–400 °C, which enhances the observed fracture. Brittle fracture occurs below the
plateau-like appearance of the curves, and the brittle-to-ductile transition temperature (BDTT) as
yielding is characterized by rather fine serrations. is exemplified by Fig. 5(a). Another specimen with
Furnace cooling from 1100 °C lowers the yield the same nominal composition at the same tem-
perature, for which an experimental error of 2 K
has to be considered, did not fracture (Fig. 5(b)).
This is believed to be due to small deviations in
composition and temperature, which indicates the
obviously rather sensitive dependence of BDTT on
composition and a rather narrow BDTT range. The
ductile fracture above the BDTT is exemplified by
Fig. 5(c). A similar behaviour was found for the
hot-rolled Fe-15%Al alloy which was recrystal-
lised at 900 °C to obtain a smaller grain size.
The observed variation of impact toughness (as
given by the impact fracture energy) with tempera-
ture is shown in Fig. 6 for various alloys. There is
indeed a sudden transition from brittle behaviour
with low toughness to ductile behaviour with high
toughness within a narrow temperature interval.
The fracture surfaces indicated transcrystalline
cleavage below the BDTT and ductile fracture
above the BDTT. This transition is shifted to
higher temperatures with increasing Al content.
The effects of prior heat treatment on toughness
Fig. 3. Compressive yield stress (10⫺4 s⫺1 compression rate) are illustrated by Fig. 7(a)–(b) for polycrystalline
as a function of temperature for various polycrystalline Fe- and monocrystalline Fe-16%Al. There is a pro-
12%Al with 50 ppm C and Fe-17%Al with less than 20 ppm C
with concurrent twinning and serrated yielding (with qualitative nounced gap between the BDTTs of the alloys with
stress-strain curves in the insets) after various treatments (Q: and without initial quench from 1100 °C, i.e. the
quench from 1100 °C, FC: furnace cooling from 1100 °C). quench shifts the BDTT to significantly higher
3236 J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242

Fig. 4. Light-optical micrograph of the microstructure of the polycrystalline Fe-17%Al alloy with less than 20 ppm C of Fig. 3,
which was deformed at 350 °C in compression after furnace cool (a) or quench (b) from 1100 °C.

temperatures which is not relaxed by subsequent BDTT observed for this alloy is 70 °C which was
anneals below the quench temperature. Only re- produced by annealing at 550 °C with subsequent
heating to the quench temperature of 1100 °C and quench. There is a steep increase of BDTT with
subsequent furnace cooling restores the state increasing anneal and quench temperature in the
before quenching and results in BDTTs in the range 550–700 °C and a distinctly reduced increase
range of those obtained without quenching. How- above this quench temperature range.
ever, the latter BDTT (quench, re-heating, furnace The increase of the BDTT with increasing Al
cooling) is still higher than that for furnace cooling content is shown in Fig. 9. The monotonous
without any quenching, and other treatments with- increase of the data for the as-cast alloys with
out quenching are in between. Deformation twins increasing temperature is interrupted by apparent
were found near the crack surfaces by metallo- relative minima at 8 and 13% Al (Fig. 9(a)), which
graphic observation in all cases, i.e. for all heat more or less correspond to the relative maxima of
treatments and fracture at room temperature and the variation of ductility with increasing Al content
above the BDTT. at room temperature [28]. Furnace cooling obvi-
The effect of annealing and quenching tempera- ously leads to higher BDTT values and the effect
ture on the BDTT is shown in Fig. 8 for as-cast seems to be more pronounced at higher Al contents
Fe-16%Al with 40 ppm C. Clearly the higher the above 12% Al (Fig. 9(a)). However, the high
BDTT is the higher the temperature of isochronal BDTT of 200 °C for the furnace-cooled alloy with
annealing with subsequent quench is. The lowest 16.5% Al and the relatively low BDTT of 75 °C
J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242 3237

Fig. 6. Impact toughness (energy absorbed in Charpy impact


test) as a function of temperature for various polycrystalline as-
cast Fe-Al alloys.

at 320 °C produced inconclusive data. Fig. 9(b)


compares the BDTT data for as-cast alloys with
the data for quenched alloys. In contrast to the as-
cast alloys, the BDTT of the alloys with quench
from 1100 °C increases monotonously with
increasing Al content. In particular, the relative
maxima for the as-cast alloys with Al contents
below 15% Al in Fig. 9(a) and (b) are avoided,
and the minima are only slightly lower than the
respective data for the quenched alloys in Fig. 9(b).
The quench from 1100 °C produces a drastic
increase of BDTT only at high Al contents above
15% Al. However, this latter increase is still within
the range of the data in Fig. 9(a).

4. Discussion

4.1. Effects of ordering

The toughness data in Fig. 7 do not indicate any


Fig. 5. Impact force-displacement curves for Charpy impact
major impact of short-range ordering on the mech-
tests of polycrystalline as-cast Fe-17%Al with brittle fracture
at 100 °C (a), bending without fracture at 100 °C (b) and ductile anical behaviour. This agrees with yield stress
fracture at 150 °C (c). behaviour at low temperatures [28]. The BDTT of
monocrystalline Fe-16%Al in Fig. 7(b) is slightly
lower than 70 °C after furnace cooling from 1100
for furnace-cooled Fe-16%Al indicate the appreci- to 600 °C with subsequent quench to room tem-
able variation of data, which may be related to the perature and slightly higher than 70 °C after fur-
different C contents of the various alloys—see the nace cooling from 1100 °C to room temperature
respective data in Fig. 9(a). An isothermal anneal whereas the BDTT of polycrystalline Fe-16%Al in
3238 J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242

Fig. 7. Impact toughness (energy absorbed in Charpy impact test) as a function of temperature for polycrystalline Fe-16%Al with
less than 20 ppm C (a) and monocrystalline Fe-16%Al with 45 ppm C after various treatments (Q: quench, FC: furnace cooling,
AC: air cooling).

Fig. 7(a) is about 80 °C after furnace cooling from the analogous conclusion with respect to the low-
1100 to 600 °C with subsequent quench to room temperature yield stress [28].
temperature and about 70 °C after furnace cooling
from 1100 °C to room temperature. However, the 4.2. Effects of quenched-in excess vacancies
prior quench from 1100 °C to room temperature
without and with various subsequent anneals at and Quenching from high temperatures produces
below 600 °C produces BDTTs above 150 °C. It is excess vacancies which act as obstacles to dislo-
concluded that only heat treatments at temperatures cation movement and contribute to hardening [28].
above 600 °C affect the mechanical behaviour of Correspondingly the highest BDTT with only a
Fe-Al alloys significantly. This again agrees with gradual transition to ductile behaviour is shown in
J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242 3239

Fig. 8. Brittle-to-ductile transition temperature BDTT as a


function of the quench temperature for polycrystalline as-cast
Fe-16%Al with 40 ppm C after anneal of 2 h at the quench
temperature and subsequent water quench (BDTTmin: minimum
BDTT observed for this alloy).

Fig. 7(a) by the quenched specimen. This quench-


ing effect on BDTT increases with increasing
quench temperature as is demonstrated by Fig. 8.
The effect becomes prevalent at Al contents above
about 13% and increases with increasing Al con-
tent as is visible in Fig. 9(b). This corresponds to
the increase of the vacancy concentration with
Fig. 9. Brittle-to-ductile transition temperature BDTT as a
increasing temperature and increasing Al content. function of Al content for polycrystalline as-cast Fe-Al alloys:
It is concluded that the observed effects of quench- (a) comparison with furnace cooled and isothermally annealed
ing on the mechanical behaviour are due to alloys (the impurity carbon contents are given by the numbers
quenched-in excess vacancies. It is emphasised that in ppm at the symbols; the C contents of the as-cast alloys
these effects are revealed most clearly only with correspond to those of the heat-treated alloys if not specified
otherwise); (b) comparison with as-cast alloys which are
the high deformation rates of impact testing. quenched from 1100 °C.
The excess vacancies get healed out only slowly
by annealings below the quench temperature with
only small effects on hardness [28]. Indeed the
BDTT is slightly reduced by anneals after quench- not have been sufficient for the complete elimin-
ing, the largest reductions resulting from the longer ation of the excess vacancies.
anneals at 300 °C and from anneals at the higher
annealing temperatures (Fig. 7(a)). These effects 4.3. Effects of Al content
are only small in view of the data for the specimens
without quench, which confirms the slow kinetics A linear increase of the yield stress with increas-
of the healing out of excess vacancies. Indeed the ing Al content was found for Fe-Al alloys at low
equilibration of the vacancy concentration was temperatures, which is produced exclusively by
reported to be very slow and complete healing out solid solution hardening [28] and which is
was reached only by cyclic annealing treatments described by Suzuki’s theory [35,36]. This harden-
[34]. This would mean that the annealing at 300 ing effect was found to be accompanied by a linear
°C of the alloy in Fig. 7(a) even for 14 days may decrease of the elongation before necking with
3240 J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242

increasing Al content, which is overlayered by few which are specific for the B2 structure and may
small deviations to higher and lower strains. not apply to the present Fe-17%Al alloy. Serrations
The toughness as characterised by the BDTT in at high temperatures were reported for partially
Fig. 9(a)–(b) shows a corresponding behaviour. D03-ordered Fe-Al with 25% Al at 520 °C for two
There is a linear increase, which is overlayered by deformation rates 4.8 × 10⫺5 and 2.4 × 10⫺4 s⫺1
few small deviations to higher and lower BDTT [38]. Again these effects were discussed with
for the as-cast alloys, with increasing Al content respect to the particular dislocation configurations
only up to about 13% Al and a noticeably steeper which are characteristic only for the D03 structure.
increase at higher Al contents. It is noticed that the In view of the high temperatures up to 590 °C for
effects of annealing at temperatures above 600 °C the serrations in Fig. 1 and the high Al content of
and subsequent quenching, which have been dis- 17%, which is in the range of beginning short-
cussed in the preceding section, occur in this Al range ordering below 600 °C, it is concluded that
content range above 13%. The deviations from lin- the observed high-temperature serrations as well as
earity for the as-cast alloys correspond to those of the yield stress drop is related to interactions
the room-temperature ductility [28], i.e. the between mobile dislocations and dynamic short-
enhanced elongation before necking and fracture range ordering, i.e. moving dislocations cut
strain at room temperature corresponds to a dynamically formed short-range order which pro-
reduced BDTT at about 8–9% Al and likewise at duces disorder and competes with thermal resto-
about 12% Al. It is concluded that both strength ration of order. Such a mechanism was designated
and ductility are controlled by solid solution hard- as pseudo-PLC effect and was modelled previously
ening with additional effects by heat treatments for long-range ordered alloys with possible appli-
above 600 °C. cation also to short-range ordered alloys [40].
The increasing Al content was found to reduce The data in Fig. 1 reveal a hardening effect of
the dislocation mobility [28]. A further evidence the accidentally dissolved small amounts of C
for this is provided by the occurrence of defor- which may also give rise to a Portevin-Le Châtelier
mation twinning, which was observed at high tem- effect. However, a previous study of serrated flow
peratures profusely for the Fe-17%Al alloy, but of a low-carbon steel, i.e. Fe with various microal-
only exceptionally for the Fe-12%Al alloy (Fig. 3). loying additions and 1300 ppm C revealed a tem-
perature range of 70–200 °C for the PLC effect at
4.4. Yield stress drop and stress-strain serrations the used deformation rate [41]. Likewise a study
of internal friction in Fe-Al-C alloys with 0–30%
Various types of yield stress drop and stress- Al and about 100 ppm C revealed relaxation peaks
strain serrations have been observed at high tem- only in the temperature range of 40–250 °C (at
peratures as is indicated in Figs. 1–3. Serrations about 1 cycle/s) [42]. A Snoek peak was reported
are visible in Fig. 1 for the Fe-17%Al alloy (with for Fe3Al at about 230 °C [43]. Thus it is con-
two different impurity contents) at high tempera- cluded that the presently observed discontinuous
tures in the range 300–600 °C. Yield stress drops flow at higher temperatures is not related to dis-
were produced by Fe-17%Al at 590 °C (Fig. 1). solved carbon. The present low carbon contents of
Such discontinuous yielding phenomena indicate 50 ppm C leads to carbide formation during slow
the Portevin-Le Châtelier (PLC) effect which cooling from high temperatures or annealings at
results from a negative strain rate sensitivity 320 °C which indicates a carbon solubility in the
because of dynamic interactions between mobile studied Fe-Al alloys of less than 50 ppm C at 320
dislocations and diffusing solutes [37]. °C [28].
The observed yield stress drop at 590 °C (Fig. In addition to the PLC effect with fine stress-
1) for Fe-17%Al is paralleled by similar phenom- strain serrations, deformation twinning occurs at
ena in partially B2-ordered Fe-Al with 25% Al at high temperatures as is indicated by the large load
temperatures at or above 550 °C [38,39]. The latter drops in Fig. 3 with audible clicks and by metallo-
were attributed to dislocation multiplication effects graphic evidence (Fig. 4). This was expected only
J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242 3241

for low temperatures [28]. Deformation twinning 앫 The brittle-to-ductile transition temperature
at high temperatures has not yet been reported for BDTT increases with increasing Al content up
disordered and ordered Fe-Al alloys [44,45]. The to 18 at.%. The BDTT is further increased by
observations in Fig. 3, which rely on metallo- quenching from high temperatures above 600
graphic evidence, indicate enhanced deformation °C only for Al contents above 13 at.% Al. This
twinning with increasing Al content in a tempera- is attributed to the presence of excess vacancies
ture range which is characteristic for the occur- which reduce the dislocation mobility at lower
rence of the high-temperature PLC effect. The temperatures.
interaction between mobile dislocations and 앫 The serrated yielding at high temperatures
dynamic short range ordering contributes to below 600 °C is attributed to the interaction of
strengthening as is indicated by the yield stress pla- mobile dislocations and short-range ordering
teau in this temperature range (Figs. 2 and 3). This and disordering reactions.
reduces the availability of mobile dislocations thus 앫 The accidentally contained small amounts of
promoting deformation twinning. However, this carbon with contents below 100 ppm, which
effect is primarily produced by the furnace-cooled contribute to hardening, do not produce ser-
alloys and only rarely by the quenched alloys rated yielding.
which also show the PLC effect. Quenching pro- 앫 Deformation twinning at high temperatures,
duces excess vacancies which reduce the dislo- which has not been reported before for Fe-Al
cation mobility at room temperature as was dis- alloys, occurs in the temperature range of high-
cussed recently [28]. Obviously the effects of prior temperature serrated yielding and is attributed
heat treatment on microstructure, dislocation distri- to the negative effects of short-range ordering
bution and short range order are also decisive for reactions on dislocation mobility.
the occurrence of deformation twinning at high
temperatures. However, the details are not clear
and need more detailed studying. Acknowledgements

The financial support by the German Bundesmi-


5. Conclusions nisterium für Bildung und Forschung (BMBF grant
no. 03N3013D) is gratefully acknowledged.
The deformation behaviour of variously treated
binary Fe-Al alloys with Al contents up to 18 at.%
and only low unavoidable impurity contents—in References
particular less than 100 ppm C—has been studied
above room temperature up to 800 °C. The follow- [1] Klöwer J. Mater Corros 1996;47:685.
ing conclusions are drawn from the results. [2] Sykes C, Bampfylde JW. JISI 1934;130:389.
[3] Ziegler N. Trans AIME 1932;100:267.
[4] Drewes EJ, Engl B, Hofmann H, Kruse J, Menne M, Reip
앫 Short-range ordering in the studied alloys, C-P, Frommeyer G, Herrmann J, Inden G, Klaus S, Sau-
which is controlled by heat treatments at tem- thoff G, Wildau M, Dannenfeldt M, Stratmann M. Höherf-
peratures below 600 °C, has no major impact ester Leichtbauwerkstoff auf der Basis von Eisen-Alu-
on the deformation behaviour. Only heat treat- minium-Legierungen. Abschlussbericht: BMBF, 2000.
[5] Drewes EJ, Frommeyer G, Stratmann M. In: For-
ments at temperatures above 600 °C affect the schungszentrum Jülich GmbH - NMT, editor. MaTech -
mechanical behaviour of Fe-Al alloys signifi- Neue Materialien für Schlüsseltechnologien des 21. Jahr-
cantly. hunderts: Jahresbericht 1999/2000, Bonn: BMBF; 2000,
앫 Quenching from high temperatures produces p. 03 N 3013.
excess vacancies which increase the brittle-to- [6] Justusson W, Zackay VF, Morgan ER. Trans ASM
1957;49:905.
ductile transition temperature BDTT. Softening [7] Morgand P, Mouturat P, Sainfort G. Acta Metall
by subsequent anneals to eliminate the excess 1968;16:867.
vacancies is a slow process. [8] Bannykh OA, Sudin IF, Kashin WI, Prokoshkin DA, Sam-
3242 J. Herrmann et al. / Acta Materialia 51 (2003) 3233–3242

arin AN. Proc Symp Metallurgy, Metallography and Phys- [26] Baker I, George EP. In: Deevi SC, Maziasz PJ, Sikka VK,
ico-Chemical Methods of Investigations, Moscow: 1963. Cahn RW, editors. International Symposium on Nickel
p. 68. and Iron Aluminides: Processing, Properties and Appli-
[9] McKamey CG. In: Stoloff NS, Sikka VK, editors. Physical cations, Materials Park: ASM Interntl; 1997. p. 145.
metallurgy and processing of intermetallic compounds. [27] Herrmann J, Inden G, Sauthoff G, Schweika W. Acta
London: Chapman & Hall; 1996. p. 351. Mater. 2003 (in preparation).
[10] Köster W, Gödecke T. Z Metallk 1982;1973:111. [28] Herrmann J, Inden G, Sauthoff G. Acta Mater. 2003; 51
[11] Leamy HJ, Gibson ED, Kayser FX. Acta Metall (in press).
1967;15:1827. [29] Herrmann J, Inden G, Sauthoff G. Acta Mater. 2003 (in
[12] Davies RG. J Phys Chem Solids 1963;24:985. preparation).
[13] Marcinkowski MJ, Taylor ME. J Mater Sci 1975;10:406. [30] Herrmann J, Inden G, Sauthoff G. Acta Mater. 2003 (in
[14] Morris DG, Munoz-Morris MA. Rev Metalurgia preparation).
2001;37:230. [31] Herrmann J, Inden G, Sauthoff G, Schweika W. In: Kopp
[15] Baker I, Nagpal P. In: Darolia R, Lewandowski JJ, Liu R, Beiss P, Herfurth K, Böhme D, Bormann R, Arzt E,
CT, Martin PL, Miracle DB, Nathal MV, editors. Struc- Riedel H, editors. Werkstoffwoche 98—Band VI: Sym-
tural intermetallics. Warrendale/PA: TMS; 1993. p. 463. posium 8: Metalle, Symposium 14: Simulation Metalle.
[16] Munroe PR. In: Deevi SC, Maziasz PJ, Sikka VK, Cahn Weinheim: Wiley-VCH; 1999. p. 389.
RW, editors. International Symposium on Nickel and Iron [32] Herrmann J. Untersuchungen zur Struktur und zum mech-
Aluminides: Processing, Properties and Applications. anischen Verhalten von Fe-reichen Fe-Al-Legierungen.
Materials Park: ASM Interntl; 1997. p. 95. Düsseldorf: VDI Verlag, 2000.
[17] George EP, Baker I. Intermetallics 1998;6:759. [33] Michels C. Dr. Ing. Dissertation, RWTH Aachen, 1994.
[18] Liu CT, George EP, Maziasz PJ, Schneibel JH. Mater Sci [34] Broska A. Dr. rer. nat. Dissertation, Universität Göt-
tingen, 1998.
Eng A-Struct Mater 1998;258:84.
[35] Neuhäuser H, Schwink C. In: Cahn RW, Haasen P,
[19] Morris DG, Morris-Munoz MA. Intermetallics
Kramer EJ, Mughrabi H, editors. Materials science and
1999;7:1121.
technology. Plastic deformation and fracture of materials,
[20] Wolff J, Franz M, Broska A, Köhler B, Hehenkamp T.
Vol. 6. Weinheim: VCH; 1993. p. 191.
In: Nathal MV, Darolia R, Liu CT, Martin PL, Miracle
[36] Suzuki H. In: Nabarro F, editor. Dislocations in solids,
DB, Wagner R, Yamaguchi M, editors. Structural Inter-
Vol. 4. Amsterdam: North-Holland Publ. Co; 1979. p. 193.
metallics 1997 (Proc. ISSI-2). Warrendale: TMS; 1997. [37] Estrin Y, Kubin LP. J Mech Behav Mater 1990;2:255.
p. 721. [38] Kettner U, Rehfeld H, Engelke C, Neuhäuser H. Intermet-
[21] Würschum R, Schaefer HE. Positron Annihilation— allics 1999;7:405.
ICPA-11—Proceedings of the 11th International Confer- [39] Schmatz DJ, Bush RH. Acta Metall 1968;16:207.
ence on Positron Annihilation 1997;255:81. [40] Brechet Y, Estrin Y. Scr Metall Mater 1994;31:185.
[22] Kim SM, Morris DG. Acta Mater 1998;46:2587. [41] Pink E, Kumar S. Mater Sci Eng A-Struct Mater
[23] Wolff J, Franz M, Broska A, Kerl R, Weinhagen M, 1995;201:58.
Köhler B, Brauer M, Faupel F, Hehenkamp T. Intermet- [42] Tanaka K. J Phys Soc Jpn 1971;30:404.
allics 1999;7:289. [43] Nagy A, Harms U, Neuhäuser H. Defects and Diffusion
[24] Sprengel W, Müller MA, Schaefer H-E. In: Westbrook JH, in Metals: An Annual Retrospective IV 2002;203-2:257.
Fleischer RL, editors. Intermetallic compounds. Progress, [44] Christian JW, Mahajan S. Prog Mater Sci 1995;39:1.
Vol. 3. Chichester: John Wiley & Sons; 2002. p. 275. [45] Yoo MH. In: Westbrook JH, Fleischer RL, editors. Inter-
[25] Stoloff NS, Liu CT, Deevi SC. Intermetallics metallic compounds—principles and practice. Progress,
2000;8:1313. Vol. 3. John Wiley & Sons; 2002. p. 403.

You might also like