You are on page 1of 7

Available online at www.sciencedirect.

com

Procedia Engineering 55 (2013) 10 – 16

6th International Conference on Cree


ep, Fatigue and Creep-Fatigue Interaction [CF-6
6]

Advances in Low-Temperrature (<0.25Tm) Creep Behavior of


o
Single and Twoo-Phase Titanium Alloys
Sreeramamurthy Ankeem∗, Zane W. Wyatt, William Joost
University of Maryland, Department of Materrials Science & Engineering, College Park, MD 20742-2115, USA

Abstract

Not too long ago, the authors found that twinning cann be a very slow process (i.e. many orders of magnitude lower than t the
speed of sound) and contribute to low-temperature (<0.25Tm) creep deformation behavior in single-phase HCP titanium
alloys. Since then, it has been shown that this time-ddependent twinning behavior can also be found in BCC titanium m alloys.
Furthermore, the time-dependent twinning was also ffound to play a significant role in the two-phase Į + ȕ titanium m alloys.
The extent of twinning and its contribution to creepp behavior was found to depend on the grain size in the single-phase
titanium alloys. In regard to single-phase beta alloys,, the stability of the beta-phase was also found to play a significant role
in twinning behavior. With respect to the two-phhase Į + ȕ titanium alloys, twinning was found to depend on the
morphology and volume fraction of the phases and sstability of the beta phase. These developments are reviewed and the
ramifications in regard to the design and selection of ttitanium alloys for various applications are presented

© 2013The
© 2013 The Authors.
Authors. Published
Published by Ltd.
by Elsevier Elsevier Ltd.under
Open access Selection and/orlicense.
CC BY-NC-ND peer-review under responsibility of the Indira
Selection and
Gandhi peer-review
Centre under responsibility
for Atomic Research. of the Indira Gandhi Centre for Atomic Research.

Keywords: Titanium, creep; twinning; mechanical propertiess; plastic deformation

1. Introduction

Single-phase and two-phase titanium (Ti) allooys are technologically important. Ti alloys are used in various
v
areas such as aerospace [1], biomedicine [2-4], naval applications[5], and energy production [6]. The su uperior
specific strength, excellent corrosion resistancee, and biocompatibility of these materials is making them m more
and more popular [7]. The uses for titanium allooys are likely to grow in the 21st century. It is therefore critical
c
to understand the microstructure and mechanicall behavior of Ti alloys.
The deformation behavior of two-phase allooys depends upon the properties of the component phases, the
morphology and volume fraction of the compoonent phases and the crystallographic relationships betweeen the
phases. When two-phase alloys are subjected too a stress, the individual phases deform differently, resullting in
elastic, elasto-plastic, and plastic deformation [8]. Therefore, deformation in these two-phase alloy ys is a
process that cannot be modeled just on individuaal phase deformation and the rules of mixing.


Corresponding Author:
E-mail address: ankem@umd.edu

1877-7058 © 2013 The Authors. Published by Elsevier Ltd. Open access under CC BY-NC-ND license.
Selection and peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.
doi:10.1016/j.proeng.2013.03.212
Sreeramamurthy Ankem et al. / Procedia Engineering 55 (2013) 10 – 16 11

One characteristic of titanium alloys is that they are prone to creep at low temperatures (<0.25Tm) and at low
stresses [9-11]. Creep can cause failure in components by a change in dimension which exceeds the allowable
tolerance[11-18]; oftentimes, components that depend heavily on titanium, such as aircraft landing gear, require
precise part tolerances so predicting low temperature creep becomes very important. Creep in α-titanium
alloys has been primarily explained by slip [15]. However, in comparison with α-titanium, studies of low
temperature creep in β and near-β alloys have been very limited with a few exceptions [19-25], and the
deformation mechanisms consist of primarily slip with some twinning in large grain sized alloys. Creep studies
of two-phase titanium alloys have primarily concerned commercial alloys [11, 12, 14-18, 26], and have been
attributed to slip in the individual phases as well as across the interphase boundaries.
Over the years, considerable research has been performed examining and seeking to improve the mechanical
properties (including low temperature creep) of two-phase alloys [27]. In many of the applications where use
of titanium is attractive, low temperature creep is an important factor, where low temperature is defined as less
than 25% of the melting temperature (1941 K for Ti). Creep is an important consideration in high-pressure
vessels, aerospace components, structural components [28] and artificial joints such as hips or knees [4]. Given
the importance of reliability in these cases, it is vital to understand the deformation mechanisms of Ti alloys
under these conditions. In this sense, the tensile strength and creep resistance of the two-phase Į + ȕ Ti alloys
are dependent upon a number of parameters such as microstructure, the properties of the component phases,
interactions between phases, and the ȕ-phase stability.
In addition to creep deformation leading to creep failure, creep deformation can also affect other failure
modes such as fatigue failure and stress corrosion cracking [9]. This is due to the fact that low temperature
creep deformation can result in such features as twins and martensite [19, 22, 29-33], which are not seen in the
respective single-phase alloys of equal composition, and which can act as fatigue crack initiation sites. To
address some of these issues, Ankem et al. have made significant advances in understanding the low
temperature creep deformation behavior of two-phase Į + ȕ titanium alloys [29, 31-37]. It has been shown that
a decrease in grain size results in a decrease in time-dependent twinning and hence a reduction in creep strain
in single phase Į and ȕ alloys [22, 29, 32], a trend opposite to that observed during high temperature creep [38,
39]. It was also shown that a fine-equiaxed microstructure is more resistant to creep strain at low temperatures
than a Widmanstätten microstructure in two phase Į + ȕ alloys, which is also contrary to observations during
high temperature creep [40]. Most recently, the authors found that an increase in the stability of the metastable
ȕ phase will result in a decrease in creep strain [41] and that whether the ȕ phase deforms by twinning or by
martensite formation depends on the volume fraction of the ȕ phase. These developments are briefly reviewed
below.

2. Single-phase alpha titanium alloys

When an Į-Ti-1.6wt% V alloy with a coarse grain size (225ȝm) was creep deformed in tension at a stress
level of 95% yield stress at 298K, time-dependent {101̄2} type twinning was observed [29, 33]. Similar
phenomena were also observed earlier in a coarse grained Į-Ti-0.4wt% Mn alloy [30]. This phenomenon is
contrary to the conventional understanding of twinning where the twin growth rates are expected to approach
the speed of sound. The activation energy for twin growth is often much less than the activation energy for
nucleation so twins have been previously thought to grow to their final lamellar thickness almost
instantaneously[42, 43].
Experiments were performed by Ankem & Oberson [33] to determine the activation energy of the creep
process from all deformation mechanisms which contribute to creep. The activation energy was found to
increase from a value of approximately 38 kJ/mol at low strain (<0.2%) to approximately 47 kJ/mol at higher
strain (•4%) [33]. This was the first time that the activation energy for creep has been calculated for an Į-
titanium alloy that deforms by twinning. The change in activation energy suggests a potential change in the
rate-controlling creep deformation mechanism, from predominately slip at the beginning of the deformation
process to a mixed mode consisting of slip and twinning [33]. This suggestion is consistent with the fact that
the value of the activation energy for deformation of Į-titanium when slip is the predominant deformation
mechanism was found to be 30 – 40 kJ/mol [33].
12 Sreeramamurthy Ankem et al. / Procedia Engineering 55 (2013) 10 – 16

The question is: why are the twin boundaries moving so slowly? It has been traditionally believed that twins
grow at speeds approaching the speed of sound in the material. To understand this phenomen non, a
crystallographic model has been developed by b Oberson & Ankem [36]. Some of the results of o this
crystallographic model are shown in Fig. 1. Thee model in Fig. 1 is an extension of the model proposed by y Song
and Gray [44, 45], where the movement of octahhedral interstitial sites are included, in addition to the mov vement
of the parent atoms. This figure shows that none of the octahedral sites are conserved, i.e. they are all
displaced. This means that if atoms such as oxyygen are present at these interstitial sites, the oxygen atomms will
interfere with the twin propagation, and the oxxygen must be diffused away from the interface either in nto the
matrix or into the newly formed twin. Thereforee, originally a solely shear and shuffle process for the form mation
of the twin is now controlled by the diffusionnal process of oxygen atoms. If indeed the oxygen difffusion
controls the twin growth, then the activation energy
e for the twin growth would correspond to the actiivation
energy of oxygen diffusion in Į-Ti. The activattion energy of twin growth in Į-Ti was calculated by meaasuring
individual twin growth rates. This activation energy
e was found to be 66 kJ/mol [33]. This activation energy
was compared to the activation energy of diffussion of oxygen in Į-Ti. At 298K, the experimentally meeasured
activation energy for oxygen in Į-titanium ranges from about 65 – 200 kJ/mol [46]. The low end of thiss range
is close to the value of the model, suggesting thhat the growth of twins is controlled by the diffusion of oxygen.
It is also possible that the slightly lower activattion energy can be related to the fact that dislocation motion or
slip is also involved in the creep deformation prrocess simultaneous with twinning and the activation enerrgy for
the slip process is lower than that for twinning.

CP lattice of Į-titanium. Projection of the lattice is onto the {112̄0} pllane. (a)
Fig. 1. Schematic illustration of {101̄2} twinning in the HC
Shear of the lattice in the <1̄101̄> direction moves only A-type Ti atoms (large, corner circles) directly to twinned positions. (b)) Arrows
give the shuffles required to move the Ti atoms in B-tyype sites (large, center circles) to their twinned B-type or C-type positions.
p
Reorientation of the lattice eliminates the octahedral sitees (small circles with “x” mark) where oxygen could reside. (c) Scchematic
illustration of a completed {101̄2} twinning event. The commpleted positions for B-type Ti atoms and the corresponding possible occtahedral
sites and the completed positions for C-type Ti atoms and the
t corresponding possible octahedral sites are indicated by the arrowss. These
models are an extension of earlier twinning models by Song S and Gray [44, 45], but now include the corresponding movem ments of
octahedral sites where oxygen may reside [36].
Sreeramamurthy Ankem et al. / Procedia Engineering 55 (2013) 10 – 16 13

3. Single-phase beta titanium alloys

Similarly to Į-titanium alloys, it was found that


t time dependent twinning and slow twin growth also plays
p a
significant role in the low temperature creep behhavior of some coarse-grained ȕ-titanium alloys. Doraisw wamy&
Ankem studied the creep behavior of a coarse-ggrained ȕ-Ti-14.8wt% V alloy with a grain size of 350 ȝm ȝ and
deformed at 95% yield stress and 298K[19]. It I was found that slip as well as {332̄}<113> twinning are a the
major deformation mechanisms,and the amounnt of twinning and hence creep strain increases with incrreasing
grain size and decreasing beta phase stability; thhe creep data was used to calculate the activation energy for the
deformation process[35]. The activation energyy was shown to increase with increasing strain, from 39 kJ/mol
to 112 kJ/mol. This variation is due to a changge in the dominant creep deformation mechanisms, from slip at
the beginning of the creep process to twinningg at the end. This suggestion is consistent with the work w of
Ramesh & Ankem [22]who have observed slip at a low strains and twinning at high strains in similar ȕ-Ti alloys.
a
Similarly to Į-titanium alloys,the slow groowth of twins was attributed to the movement of octaahedral
interstitial oxygen atoms away from the twiin-matrix interface. Detailed crystallographic modelss were
developed for the {332̄ }<113> twinning process which involves both shear and shuffles as well as the
destruction of the octahedral interstitial sites at a the twin-matrix interface [35]. Some of the details of this
crystallographic model are shown in Fig. 2. Itt is suggested that twinning shear alone does not create a valid
twin and an additional shuffle is required to move Ti atoms to their final positions for this particular twin n. The
reorientation of the atoms eliminates the spacce where the interstitial oxygen atoms can reside (octaahedral
interstitial site), and they are required to diffuse to an acceptable site. Based on the measured twin growtth rates
in the ȕ alloy, the activation energy for the creeep deformation process was calculated as 105 kJ/mol [35]]. This
is close to the value for the diffusion of oxygenn in ȕ-Ti of 122 kJ/mol determined by Claisse and Koenig [47].
These results suggest oxygen atoms play a significant role in controlling the growth of twins; thereby
controlling the extent of creep strain.

Fig. 2. Projection of BCC substitutional and octahedral inteerstitial atoms onto the (11̄0) plane. (a) Untwinned structure showing diirection
of [113] twinning shear, and (b) Final twinned sttructure showing mirror symmetry across twin/matrix interface [35].

4. Two-phase alpha + beta titanium alloys

In regard to two-phase Į + ȕ titanium alloyss, it was found that the deformation mechanisms could be quite
different from those observed in the respective single-phase
s Į and ȕ alloys. For example, in the case of a 51%Į
– 49% ȕ Ti-8.1wt% V alloy, with Widmanstätteen Į+ ȕ microstructure, the creep deformation mechanism ms were
found to be twinning in the Į phase and stress-iinduced martensite in the ȕ phase [32, 34]. There were at a least
two surprising results here. First, only slip waas found to be a deformation mechanisms in single phasee Į-Ti-
1.6wt% V where the grain size was small (<60ȝ ȝm) [29]; however, in the case of this two phase alloy, tw
winning
was observed in the Į phase even though the Į plate thickness is about 3.0ȝm. Secondly, in the case of the
14 Sreeramamurthy Ankem et al. / Procedia Engineering 55 (2013) 10 – 16

single-phase ȕ alloy discussed in Section 3, twiinning along with slip were found to be the major mechaanisms;
however, in this two-phase alloy, stress induuced martensite was found to be a major creep deform mation
mechanism in the ȕ phase. It is to be noted thhat the chemical composition of the beta phase in both single
phase and two phase alloys is the same. Additionally, twinning can occur in small equiaxed Į grains in a two-
phase Ti-8.1 wt% V alloy, suggesting that the interactions seen between phases with Widman nstätten
microstructures also play a role in equiaxed miccrostructures[32]. These interesting results have been exp
plained
on the basis of elastic interaction stresses [27, 32,
3 34], which can oftentimes be much higher than the applied
a
stress, triggering additional mechanisms.
During high temperature creep deformationn, a Widmanstätten microstructure is known to have a greater g
resistance to creep strain due to the orientation relationship which exists between the Į and ȕ phases [338-40].
This platelet structure hinders interphase interfface sliding which is a primary deformation mechanism during
high temperature creep. However, as Figuree 3 shows, at low temperatures (298 K), a Widman nstätten
microstructure exhibits about 13% greater creepp strain than an equiaxed microstructure for an equivalent testing
time. Figure 4 demonstrates the reason for thhis unexpected trend [32]. As evidence by the figure, coarse
deformation products (specifically twins in the Į phase and stress induced martensite in the ȕ phase) caan span
across phase boundaries and across large distannces. It is the same Burgers orientation relationship that helps
reduce creep strain during high temperature creep c that provides an easy pathway for the transmisssion of
deformation mechanisms between phases duringg low temperature creep, thus increasing creep strain as op pposed
to an equiaxed microstructure. Whether marteensite instead of twinning occurs in two-phase alloys waas also
found to depend on the volume fraction of phasees.

Fig. 3. Creep curves of Ti-8.1 wt% V alloys, tested at 298 K and 95% YS. Note the increased
creep strain of the Widmanstätten microostructured alloy over a similarly grain sized equiaxed alloy.

Fig. 4. (a) TEM micrograph and (b) SEM image showingg connection between twinning in Į-phase and stress induced martensitee in ȕ-
n [32].
phase. Note the connected deformation features can span many grains and cover large distances, leading to increased creep strain
Sreeramamurthy Ankem et al. / Procedia Engineering 55 (2013) 10 – 16 15

5. Summary

Given that the diffusion of oxygen, and hence the rate of twin growth in Į and ȕ titanium alloys, depends on
the number of interstitial sites occupied by oxygen, these results suggest a variation in the amount of oxygen in
the alloy can have a significant effect on the amount of room temperature creep deformation. Further, the
twin/matrix boundary can act as a nucleation site for fatigue cracks and other deformation mechanisms which
would have another direct effect on the structural integrity of titanium components. Since this behavior is
similar in both HCP and BCC crystal structures, this suggests that the models presented here are broadly
applicable to describe the effect of interstitial impurities on twinning during creep of many crystalline materials
such as metals, intermetallics, ceramics, and high-temperature super-conductors with a variety of crystal
structures; where impurities can play roles in deformation mechanisms.
In two-phase Į + ȕ Ti alloys, many factors contribute to the possible deformation mechanisms during low
temperature creep; including grain size, ȕ-phase stability, elastic interaction stresses, and crystallographic phase
orientation relationships. The mechanisms in the two-phase Į +ȕ alloys included slip and twinning in the Į-
phase, and slip, stress induced martensite, and twinning in the ȕ-phase. It was shown that new deformation
mechanisms not seen in the respective single-phase alloys, such as stress induced martensite, can be explained
by the interaction stresses between the two phases and the increased creep strain of Widmanstätten
microstructures can be explained by these new deformation mechanisms as well as by the orientation
relationship which exists between the two phases. In the future, the most important factor to predict and design
multi-phase alloys for increased creep resistance, will be the control of the interactions between phases, which
in turn depend on all of the factors discussed.

Acknowledgement

This work was supported by the National Science Foundation under Grant Number DMR-0906994.

References
[1] F-22 Raptor Team Website. © Lockheed Martin Aeronautics Company 2009; Available from: http://www.f22-raptor.com/.
[2] T. Akahori, et al., Materials Science & Engineering C-Biomimetic and Supramolecular Systems,25 (2005) 248-254.
[3] S.C.P. Cachinho and R.N. Correia, Journal of Materials Science-Materials in Medicine,19 (2008) 451-457.
[4] M. Geetha, et al., Progress in Materials Science,54 (2009) 397-425.
[5] B.W. Neuberger, et al., Metallurgical and Materials Transactions a-Physical Metallurgy and Materials Science,42A (2011) 1296-
1309.
[6] K.L. Murty. Effects of Alloying and Cold-Work on Anisotropic Biaxial Creep of Titanium Tubing. in Proceedings of the Tenth World
Conference on Titanium. 2003. Hamburg, Germany.
[7] E.W. Collings, The Physical Metallurgy of Titanium Alloys, American Society for Metals, Metals Park, OH (1984), p.
[8] S. Ankem and H. Margolin, Metallurgical Transactions a-Physical Metallurgy and Materials Science,17 (1986) 2209-2226.
[9] H. Andenstedt, Metal Progress,56 (1949) 658.
[10] D.R. Luster, Wentz, W.W., Kaufman, D.W., Materials and Methods,37 (1953) 100.
[11] A.W. Thompson and B.C. Odegard, Metallurgical Transactions,4 (1973) 899-908.
[12] H.P. Chu, Journal of Materials,5 (1970) 633.
[13] M.A. Imam and C.M. Gilmore, Metallurgical Transactions a-Physical Metallurgy and Materials Science,10 (1979) 419-425.
[14] W.H. Miller, et al., Metallurgical Transactions a-Physical Metallurgy and Materials Science,18 (1987) 1451-1468.
[15] T. Neeraj, et al., Acta Materialia,48 (2000) 1225-1238.
[16] B.C. Odegard and A.W. Thompson, Metallurgical Transactions,5 (1974) 1207-1213.
[17] S. Suri, et al., Mat Sci Eng a-Struct,234 (1997) 996-999.
[18] S. Suri, et al., Acta Materialia,47 (1999) 1019-1034.
[19] D. Doraiswamy and S. Ankem, Acta Materialia,51 (2003) 1607-1619.
[20] C.A. Greene, Ankem, S., Ambient Temperature Tensile and Creep Deformation Behavior of Alpha and Beta Titanium Alloys, TMS,
(1993), p. 309-319.
[21] C.A. Greene, Fundamental Studies on the Ambient Temperature Creep Deformation Behavior of Alpha, Beta, and Alpha-Beta
Titanium Alloys, in Department of Materials Science and Engineering. 1994, University of Maryland, College Park: College Park,
MD. p. 92.
[22] A. Ramesh and S. Ankem, Metallurgical and Mater. Trans. a-Physical Metallurgy and Materials Science,33 (2002) 1137-1144.
16 Sreeramamurthy Ankem et al. / Procedia Engineering 55 (2013) 10 – 16

[23] C. Huang and M.H. Loretto, Mat Sci Eng a-Struct,193 (1995) 722-728.
[24] L. Ponsonnet, et al., Mat Sci Eng a-Struct,262 (1999) 50-63.
[25] G.B. Viswanathan, et al., Acta Materialia,50 (2002) 4965-4980.
[26] S. Ankem and H. Margolin, Journal of Metals,32 (1980) 38-38.
[27] S. Ankem, et al., Progress in Materials Science,51 (2006) 632-709.
[28] B.W. Neuberger, Greene, C., Gute, G., Mat. Res. Soc. Symp. Proc.,716 (2002) JJ11.7.1-JJ11.7.8.
[29] A.K. Aiyangar, et al., Metallurgical and Materials Transactions a-Physical Metallurgy and Materials Science,36A (2005) 637-644.
[30] S. Ankem, et al., Scripta Metallurgica Et Materialia,30 (1994) 803-808.
[31] A. Jaworski and S. Ankem, Reviews on Advanced Materials Science,10 (2005) 11-20.
[32] A. Jaworski and S. Ankem, Metallurgical and Materials Transactions a-Physical Metallurgy and Materials Science,37A (2006)
2755-2765.
[33] P.G. Oberson and S. Ankem, International Journal of Plasticity,25 (2009) 881-900.
[34] A. Jaworski and S. Ankem, Metallurgical and Materials Transactions a-Physical Metallurgy and Materials Science,37A (2006)
2739-2754.
[35] P.G. Oberson and S. Ankem, Physical Review Letters,95 (2005) -.
[36] P.G. Oberson, et al., Scripta Materialia,65 (2011) 638-641.
[37] Z. Wyatt and S. Ankem, Materials Science Forum,654-656 (2010) 863-866.
[38] F. Appel, et al., Intermetallics,8 (2000) 1283-1312.
[39] M.J.R. Barboza, et al., Mat Sci Eng a-Struct,428 (2006) 319-326.
[40] B.D. Worth, et al., Metallurgical and Materials Transactions a-Physical Metallurgy and Materials Science,26 (1995) 2947-2959.
[41] Z. Wyatt and S. Ankem, J Mater Sci,45 (2010) 5022-5031.
[42] J.W. Christian and S. Mahajan, Progress in Materials Science,39 (1995) 1-157.
[43] P. Gumbsch and H. Gao, Science,283 (1999) 965-968.
[44] S.G. Song and G.T. Gray, Acta Metallurgica Et Materialia,43 (1995) 2325-2337.
[45] S.G. Song and G.T. Gray, Acta Metallurgica Et Materialia,43 (1995) 2339-2350.
[46] Z. Liu and G. Welsch, Metallurgical Transactions a-Physical Metallurgy and Materials Science,19 (1988) 1121-1124.
[47] F. Claisse and H.P. Koenig, Acta Metall Mater,4 (1956) 650-654.

You might also like