14
7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–… http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 1/14

The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

Embed Size (px)

Citation preview

Page 1: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 1/14

Page 2: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 2/14

a range of alloys. However, attempts at providing a quan-titative explanation of the strengthening effect of clustersand co-clusters have been limited and have been based ona limited physical basis. In this paper we will present a the-ory for the thermodynamics of and the strengthening dueto co-clusters in metallic alloys; thus providing the theoret-ical basis for this strengthening mode in metallic alloys.The model is applied to Al–Cu–Mg alloys and testedagainst an extensive amount of published and new data.This alloy system is chosen for the interlinked reasons of alloy availability, data availability, relative ease of samplepreparation and analysis, and importance in engineeringapplications. Al–Cu–Mg alloys are extensively used forlightweight, fatigue resistant structures such as lower wingskins and fuselage of airplanes [8,9] as well as a range of other applications. The general validity of the theory isthought to not be limited to the investigated alloys: thedeveloped theory will indicate that co-cluster formationshould occur in many supersaturated alloys.

The method of choice for identifying thermodynamics of the co-clusters is calorimetry, as through calorimetry boththe enthalpy changes of the system and the metastable solvican be determined. However, calorimetry, in isolation, can-not identify clusters or phases. Indications for the existenceof co-clusters have mostly been obtained with 3 dimen-

sional atom probe (3DAP) (and more recently supported

by Nuclear Magnetic Resonance (NMR) [10,11], X-rayabsorption spectroscopy (XAS) [12] and positron annihila-tion lifetime spectroscopy (PALS) studies [13]).

1.2. Co-clusters in Al–Cu–Mg alloys

For solution treated Al–Cu–Mg based alloys with com-position in thea+S phase field, room temperature hardeningoccurs within several days (several weeks if the Cu content isvery low) and hardening during artificial ageing is very fast,typically within minutes [14–16]. Whilst diverging explana-tions have been proposed for this phenomenon, recentexperiments in combination with careful reassessment of existing evidence has indicated that the phenomenon iscaused by Cu–Mg co-clusters [7]. In assessing the experimen-tal evidence, we should note that immediately after the rapidhardening and during the room temperature hardening inthese Al–Cu–Mg based alloys no distinct precipitate phasecan be detected by conventional TEM [17], but differential

scanning calorimetry (DSC) experiments clearly show achange in the free energy of the alloys indicating that a sig-nificant fraction of the dissolved atoms have been involvedin a reaction [18,19]. There are no reports of any orderedstructures at this stage. Results from several atom probefield ion microscopy (APFIM) and three-dimensional atomprobe (3DAP) studies reveal Cu–Mg co-clusters of typically1 nm diameter (10–40 atoms) [7,20,21] which by someauthors are held directly responsible for the rapid hardeningreaction [20,21]. Information on the composition of Cu–Mgco-clusters is limited. Ringer et al. [20,21] observed that in ahigh purity Al–1.7at.%Mg–1.1at.%Cu aged for 5 min at

150 °C clusters with high Cu:Mg ratios existed, whilst Star-ink et al. [7] found that in an Al–2.78 wt%Cu–0.44 wt%Mgand in an Al–2.78 wt%Cu–1.05 wt%Mg (both alloys arecommercial purity with Mn additions) the Mg/Cu ratio of the clusters varied from one cluster to another. They alsocontain vacancies [17].

Due to the discovery of rapid hardening Al–Cu–Mgbased alloys early in the 20th century, and its subsequentextensive application, the reaction has been extensivelystudied and it can be considered one of the classic issuesin metallurgy. But early researchers did not have the benefitof HREM and atom probe analysis, and some unsubstanti-ated assumptions have entered the literature. Especially inpre-1990s work early hardening reaction in Al–Cu–Mgalloys was attributed to the formation of Cu and Mg con-taining structures which were termed Guinier-Preston-Bag-aryatski (GPB) zones [22]. This term was coined in the1950s and has been used for decades without evidence of what they may consist of. Consequently a range of oftenconflicting interpretations of these GPB zones haveappeared in the literature. Only very recently a clearer pic-ture of GPB zones using scanning transmission electronmicroscopy (STEM) supported by first principle modellingof the structures is starting to emerge [23]. It is the opinionof the authors that due to differing interpretation of its

meaning and the lack of experimental evidence, the term

Atom A (Mg)

Atom B (Cu)

AB = 2 atom co-cluster= dimer

AB + AB ( π /3) 4 atom co-cluster

AB + AB ( π /3) + AB (π /3) co-cluster

AB + AB (0) 4 atom co-cluster

AB + AB (π /3) + AB (-π /3) co-cluster

Fig. 1. A 111 plane in an FCC lattice in a ternary alloy containing co-clusters, with a notation describing the co-cluster constellation.

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2377

Page 3: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 3/14

‘GPB zone’, as related to the rapid hardening in Al–Cu–Mgbased alloys, has become an obstacle to the proper under-standing of the mechanisms for rapid hardening. Hencewe will here avoid using the term as much as possible.

2. A model for the thermodynamics of and strengthening due

to co-clusters

 2.1. The thermodynamics of co-clusters

For modelling the thermodynamics of the co-clustersone may consider the possibility of the co-clusters forminga range of possible constellations in the FCC lattice. Froma virtually unlimited number of possible constellations,some possible constellations that are thought to be energet-ically favoured in Al–Cu–Mg alloys (see Section 6) areillustrated in Fig. 1. The smallest co-cluster is a dimer of nearest neighbours of dissimilar alloying atoms Fig. 1a.

In considering an approach to modelling the thermody-namics, we considered three approaches to modelling inorder of increasing complexity:

(I) A regular solution model.(II) A quasi-chemical model which takes nearest neigh-

bour interactions into account (see e.g. [24]).(III) A first principles total energy calculation (FP-CVM)

(see for instance [25]).

As described below, a dimer treated in a regular solutionmodel (with a single bond energy), was found to fullyexplain all of a very extensive set of calorimetry data on

Al–Cu–Mg alloys. Hence we will fully describe this treat-ment, and limit our consideration of more complex treat-ments to a brief discussion at the end of this subsection.

Within the regular solution model, the solvi of thephases and co-clusters in the ternary system can be derivedas follows. Consider alloying elements A and B in a hostmetal M, the total number of the respective elements inthe system are N A, N B and N M. The Gibbs free energy,G , of the system can be approximated as:

G ¼  H o À N clD H A – B À TS  ð1Þwhere N cl is the number of A–B dimers, H o is a constant

reference enthalpy andD

H A–B is the enthalpy of formationof an A–B dimer from the random solution, i.e. theenthalpy of the reaction in which one A atom and one Batom, originally in random solution, form an A–B dimer.The entropy of the system is proportional to the logarithmof the number of states w in the system:

S ¼k B lnw¼k B lnN !

ð N M!ð N A À N clÞ!ð N B À N clÞ!Þ12 N cl!

ð2Þ

where k B is Boltzmann’s constant. For equilibrium itholds oG /oN cl = 0. Performing this derivation using thelatter two equations (using the Stirling approximation

lnN ! = N lnN À N ) then provides that at equilibrium:

cAcB ¼ expðÀ2Þ expÀD H A – B

 RT 

ð3Þ

where cA is the solubility (i.e. the equilibrium molar frac-tion) of element A in the M-rich host lattice, cB is the sol-ubility of element B in the M-rich host lattice. It can benoted that this corresponds to the more generally applied

‘solubility product’ equation:

cAcB ¼ C 2 expÀD H A – B

 RT 

ð4Þ

where C  is a constant. In the latter form, no attempt ismade to fully analyse the entropy term, and only the entro-py of the atoms belonging to the Al-rich phase is consid-ered. Thus constant C , which is related, amongst others,to the entropy of the clusters or precipitates, is considereda fittable parameter. In a more general form, the solubilitylimits of co-clusters according to the regular solution of any complex MmAaBb is given as:

c Aac Bb ¼ C 2 exp

ÀD H MmAaBb

 RT 

ð5Þ

where D H MmAaBb is the enthalpy of formation of theMmAaBb complex, i.e. the enthalpy of the reaction in whicha atoms of type A, b atoms of type B and m atoms of typeM, originally part of the M-rich phase, form an MmAaBb

complex, which can be an ordered or disordered clusteror precipitate.

The latter two expressions can be tested through mea-surement of the heat evolved in the reactions of formationor dissolution of the clusters or precipitates. Particularly,when a calorimetry experiment is conducted such that the

total heat evolved on completing of the co-clustering reac-tion is measured, exothermic heat evolved, DQ, should thenbe given by:

DQ ¼ ð xoA À xAÞa

D H MmAaBb ð6Þ

where xoA is the composition of the M-rich phase at the startof the reaction, and xA is the composition of the M-richphase on completion of the reaction. xA can be solved byconsidering that the composition during the reaction upto completion of the reaction is given by:

 xoA À xA ¼ abð xoB À xBÞ ð7Þ

whilst at completion:

c Aac Bb ¼  x Aa x Bbðat completion of reactionÞ ð8Þ

Thus xA can be solved from Eqs. (5), (7), and (8). (Forparticular values of the ratio a:b, the equations can besolved analytically, particularly a = b = 1 simplifies theequations considerably, but in general a simple iterativescheme is needed.)

An alternative treatment suggested in the literature [24]involves a quasi-chemical model which takes nearest neigh-bour interactions into account. Analysis of this model (seeSection 6) showed that it cannot explain all the experimen-

tal observations.

2378 M.J. Starink, S.C. Wang / Acta Materialia 57 (2009) 2376–2389

Page 4: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 4/14

Solution treated commercial alloys typically containsome small amounts of undissolved intermetallic phasesor dispersoids, which reduce the amount of solute atomsavailable for co-cluster formation and subsequent precipi-tation. The composition of the supersaturated fcc phaseafter solution treatment can generally be predicted from

thermodynamic models (e.g. phase diagrams of the alloy).In commercial and commercial purity Al–Cu–Mg basedalloys, x-Al7Cu2Fe, S-Al2CuMg and T-Al20Cu2Mn3 canbe present [26,27]. A method for calculating the amountsof these phases is described in the Appendix. This methodis applied for all model predictions.

3. Strengthening due to co-clusters

Clusters and co-clusters are shearable and hence theOrowan strengthening mechanism is not applicable. Sev-eral strengthening mechanisms related to obstacle shearingneed to be considered:

(a) Order strengthening (including stacking faultstrengthening) or configurational strengthening.

(b) Modulus hardening.(c) Chemical hardening.

3.1. Order strengthening 

The strengthening due to an ordered phase which con-tains short or long range order depends on the way the dis-locations interact with the precipitate or clusters. If longrange order is present, dislocations can travel in groups

due to the anti phase boundary (APB) [28–32]. For thepresent case of co-cluster strengthening, the order can beconsidered to be extremely short ranged and as restorationof order by dislocations travelling in groups is not possible,dislocations will travel alone. The work done in deformingthe lattice through movement of dislocations hampered by(co-)clusters equals the change in energy related to theshort range order per unit area on slip planes, i.e. [33,34]:

DsSRO ¼ cSRO

bð9Þ

where csro is the change in energy per unit area on slipplanes (which in most cases are {111} planes) on the pass-

ing of one dislocation (a.k.a. the diffuse interphase bound-ary energy [33,35]). In the analysis of csro we could attemptto include several orders of near neighbour interaction forinstance through considering a cluster variation approach[36]. However, we will here use a simplified approach thatconsiders only nearest neighbour interactions. This is donefor the following reasons: (i) we do not have sufficient accu-rate data to determine all the k th interaction parameters,and (ii) the simpler nearest neighbour method provides amore intuitively acceptable analysis, which is a benefit asseveral competing analysis of strengthening due to co-clus-ters have been accepted. We will consider here co-clusters

that consist of a single pair of atoms A and B. The amount

of A atoms in the co-clusters is yA, the amount of B atomsin the co-clusters is yB. The amount of A atoms in the M-rich phase is xA, the amount of B atoms in the M-richphase is xB. In the (M-rich) fcc lattice structure each atomhas 12 nearest neighbours and each atom adjacent to a slipplane will have 3 nearest neighbours on the other side of 

the {11 1} slip plane. The area density of A–B nearestneighbour bonds crossing the slip plane is:

qA – Bðnd ¼ 0Þ ¼ ð y A þ y BÞS 111

¼ 4ð y A þ y BÞ ffiffiffi3

p b2

ð10Þ

where S 111 is the area on the 111 plane occupied by oneatom. On passing of one dislocation through the AB co-cluster a part of the A–B bonds present before the passingwill be destroyed and some will be retained (see Fig. 2).From the geometry of the fcc lattice it follows that for asingle dislocation on the 111 plane, 1/3 of the A–B nearestneighbour pairs will remain nearest neighbours. The pass-ing through of a dislocation can also create A–B nearest

neighbour pairs. The area density is given by 2/3 yAxB +xAxB + 2/3 yBxA + xAxB = 2/3 yAxB + 2/3 yBxA + 2xAxB.Thus the total area density of A–B nearest neighbourbonds crossing the slip plane after passage of one disloca-tion is:

Atom A (Mg)

Atom B (Cu)

shear plane

shear plane

shear plane

Fig. 2. A 1 1 1 plane in an FCC lattice with a 2-atom co-cluster being cutby a dislocation. Top shows before and after with the co-cluster remainingintact in a rotated form; middle shows before and after with the co-clusterbeing eliminated, which requires an energy input; and bottom showsbefore and after in the case where the passing of one dislocation creates a

co-cluster, which releases energy.

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2379

Page 5: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 5/14

qA – Bðnd ¼ 1Þ¼ 4 ffiffiffi3

p b2

2

3ð y A þ y BÞþ2

3 y A xB þ 2

3 y B xA þ 2 xA xB

ð11ÞThe change in area density of A–B nearest neighbourbonds crossing the slip plane on passing of one dislocationis

qA – Bðnd ¼ 0Þ À qA – Bðnd ¼ 1Þ

¼ 4 ffiffiffi3

p b2

2

3ð y A þ y BÞ À 2

3 y A xB þ 2

3 y B xA þ 2 xA xB

ffi 4 ffiffiffi3

p b2

2

3ð y A þ y BÞ

ð12Þwhere the approximation is valid for dilute alloys. Passageof further dislocations will cause further changes in areadensity of A–B nearest neighbour bonds crossing the slipplane. As the magnitude of these changes will decrease

on passing of each further dislocation, the first dislocationexperiences the greatest resistance. Considering in approx-imation that the energy related to co-clusters is defined bythe enthalpy of the nearest neighbour bond DH A–B we canapproximate:

DsSRO ¼cSRO

b¼D H A – B

b½qA – Bðnd ¼0ÞÀqA – Bðnd ¼1Þ

¼D H A – B

b3Â 4 ffiffiffi

3p  2

3ð y A þ y BÞÀ 2

3 y A xBþ2

3 y B xA þ2 xA xB

ð13Þ

3.2. Modulus hardening 

Modulus hardening is a strengthening mechanism that isdifficult to deal with theoretically [28], leading to quitecomplex expressions. The very much simplified treatmentby Cartaud et al. [37] (see also Ref. [38]), which has beenadopted in a range of works, approximates the strengthen-ing due to difference in shear modulus as:

Dsm ¼ Dl

4p ffiffiffi

2p 

 ffiffiffiffiffiffi f cl

p ð14Þ

where f cl is the volume fraction of the clusters and Dl isthe difference in shear modulus between zones, l

cl, and sur-

rounding metallic phase. In the works that apply the latterequation, lcl has invariably been determined by fitting tostrength data. A more realistic way of obtaining lcl is byapproximating it as a weighted average of the moduli of the individual pure substances, i.e. for an MmAaBb cluster:

lcl ¼ mlM þ alA þ blB

mþ aþ b ð15Þ

For example, for larger clusters in Al–Cu–Mg, where Cu,Mg and Al content may be tending towards the composi-tion of S (i.e. 25%Cu, 25%Mg and 50%Al) the above equa-tion gives an estimation of lcl, of 30 GPa (using lAl = 26.2,

lCu = 48.3 and lMg = 17.3 GPa from [39]).

A more detailed approximation of the modulus harden-ing can be obtained by a calculation of the interactionforces of screw and edge dislocations (including the maxi-mum interaction force F o) as they pass through the area(precipitate) with differing modulus. This type of calcula-tion has been performed by Nembach [40] for a spherical

precipitate and various models of the dislocation. Basedon Friedel’s approximation:

Ds ¼ 1:4F 3=2

o f 1=2

rbð2pS Þ1=2ð16Þ

Nembach’s analysis [40] showed that the critical shearstress due to modulus hardening can be represented as:

Dsm ¼ 1:4a3=2Dl3=2 f 1=2b

r p

b

3b=2À1

ð2pS ÞÀ1=2 ð17Þ

where S  is the dislocation line tension, and a and b areparameters that are obtained from the dislocation precipi-tate interaction models. As different dislocation core mod-

els predict somewhat different interaction forces betweendislocation and precipitate, some uncertainty in Dsm re-sults. Averaged values are obtained for a = 0.096 andb = 0.76, respectively.

3.3. Chemical hardening 

Chemical hardening, in its classical meaning, is due tothe interfacial energy between precipitate and matrix, andstrength increments a linearly proportional to this interfa-cial energy. It is conceptually difficult to relate a cluster(even a dimer) to an interfacial energy. As an estimate of 

Dsch we considered the classical expression for chemicalhardening and the value for the interfacial energy providedin [7]. This predicts Dsch $ 0.03Dssro, i.e. Dsch is virtuallynegligible.

4. Experimental

The compositions of the alloys studied are presented inTable 1. Alloy 1 was produced as sheet through casting,homogenising, hot rolling in several passes, solution treat-ment at 500 °C and subsequent cold deformation throughcold rolling to 5% or 10%. This alloy was aged for 2, 7

or 21 days at 170 °C.Alloy 2a and 2b are different batches of plate with iden-

tical target composition. Alloy 2a was solution heat treatedat 550 °C for 30 min, and subsequently quenched. Ageingwas carried out at 180 °C for a range of times between 1and 8 h. DSC samples were prepared from the heat treatedsamples. Alloy 2b was produced through casting, followedby hot and cold rolling (see [41,42]). DSC samples of alloy2b were resolution treated at 460 °C and aged at 140 and160 °C.

Alloys 3 and 4 were produced as hot rolled plate. DSCsamples were re-solution treated at 495 °C for 30 min,

quenched into room temperature water, weighed and intro-

2380 M.J. Starink, S.C. Wang / Acta Materialia 57 (2009) 2376–2389

Page 6: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 6/14

duced into the calorimeter. DSC runs started 3 min afterthe quenching.

All DSC samples were disc shaped (5 mm diameter,approximately 1 mm thick). DSC measurements were per-formed using a Perkin–Elmer calorimeter Pyris 1 DSC,with pure nitrogen purge gas as protective gas atmosphere.Heating rates from 2.5 to 150 °C/min was used. The calo-rimeter was calibrated with an Indium standard at all blocktemperatures used (À30, À50, À100 and À140 °C). AllDSC traces were corrected by subtracting a baseline

obtained from a DSC run with empty aluminium pans.Further corrections for imperfections of the DSC as wellas correction for the heat capacity contribution not relatedto reactions of the samples is needed [43]. The optimumchoice for this correction is alloy dependent and was con-sidered in [44]. For the Cu-lean alloys 1 and 2, further cor-rection was achieved by subtracting a linear function of  T 

fitted to the initial part of the curve, where no reactions arethought to occur. Alloys 3 and 4 were further corrected bya second order polynomial which described the combinedeffect of heat capacity difference of the sample and refer-ence and baseline fluctuations. The coefficients of the sec-

ond order polynomial were fitted by assuming that theheat effects due to reactions were zero at the following threepoints: (1) at the initial part of the curves where no reactionoccurs; (2) at the completion of the S dissolution effect; and(3) at the point where the precipitation effect of the clustersends and the dissolution effect of the clusters starts. Thusall DSC curves presented represent heat flows caused byreactions, which are directly related to enthalpy changes:

D H ¼Z T 2T 1

cpdT  ð18Þ

where cp is the heat flow measured by the DSC (the heatcapacity at constant pressure) between the selected temper-

atures T 1 and T 2 (selected DSC data on alloys 3 and 4 istaken from [45]).

Possible experimental error in the temperature determi-nation in DSC experiments is mainly due to sample vari-ability and instrument calibration. The overall accuracy istypically about 1 to 2 °C. The main source of inaccuracyfor the determination of the part of DH  due to reactionsis the choice of the points in the baseline correction andto a lesser extend sample variability and instrument cali-bration. Typical accuracies are in the order of 3% or0.2 J/g, whichever is the higher.

Course intermetallic phases (typically larger than

100 nm) in Alloys 1–4 were studied in a Jeol 6500F Field

Emission Gun Scanning Electron Microscope (FEG-SEM) equipped with Energy Dispersive Spectrometer(EDS) using mechanically ground and polished samples.

5. Results and analysis

5.1. Thermodynamic data

As a result of the very low Cu content, DSC thermo-grams of alloy 1 generally showed very weak thermal

effects. Samples aged for 2 days at 170°C show a smallendothermic effect which peaks at 220 °C (curves not pre-sented). Due to the very low heat flows and slight baselineinstability the DSC curves of alloy 2 aged at 180 °C (Fig. 3)become progressively more erratic beyond about 250 °C.However up to about 250 °C the DSC thermograms of alloy 2 are consistent and when aged for 4 to 48 h at180 °C the alloy, showed a substantial endothermic effectpeaking at 230 °C. This effect is seen to develop graduallyduring ageing for 1, 4 and 8 h, reaching a maximum heatcontent after 8 h. This gradual development of the endo-thermic effect relates to a gradual increase in G of the sam-

ples on ageing. TEM data on the same alloy shows thatafter 4 and 8 h of ageing, coherent nanometer-sized needleand lath shaped particles are present [46,47]. After 1 h age-ing these precipitates could not be detected in the TEM. Atthis stage of the ageing the DSC thermograms show a dis-solution effect that peaks at about 210 °C, with a heat con-tent that is substantially lower than that obtained after 4and 8 h.

Table 1Chemical composition of alloys (at%).

Alloy Cu Mg Mn Si Fe Cu/Mg

Alloy 1: Al–1.3Mg–0.07Cu–0.4Mn 0.07 1.34 0.44 0.19 0.19 0.05Alloy 2a: Al–3Mg–0.2Cu–0.1Mn 0.18 3.16 0.12 0.12 0.10 0.06Alloy 2b: Al–3Mg–0.2Cu–0.1Mn 0.17 3.09 0.12 0.14 0.09 0.06Alloy 3: Al–1.2Mg–1.2Cu–0.2Mn 1.19 1.19 0.20 0.03 0.02 1.00

Alloy 4: Al–1.5Mg–1.8Cu–0.2Mn 1.82 1.49 0.20 0.02 0.02 1.22

-0.01

0

0.01

0.02

0 100 200 300 400

T (°C)

   H  e  a   t   F   l  o  w   (   W   /  g   )

J366-0h

J366-1h

J366-4h

J366-8h

J366-24h

J366-48h

0 h

24 h48 h

1 h8 h 4 h

Fig. 3. DSC thermograms of solution treated alloy 2 Al–3.2Mg–0.18Cu

after ageing for 0, 1, 4, 8, 24 and 48 h at 180 °C.

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2381

Page 7: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 7/14

DSC curves of the solution treated alloy 2B aged at140 °C are presented in Fig. 4. This data shows that in asquenched samples an exothermic reaction occurs at about120–180 °C. When aged at 140 °C this exothermic reactiondisappears and the heat evolved in the endothermic reac-tion between 220 and 270 °C increases. The ageing treat-

ments considered here are less severe than the 4 and 8 hof ageing at 180 °C in which ordered nano-scale precipi-tates were detected in alloy 2. Thus, these precipitatesshould not form during ageing at 140 °C. The effectsobserved in Fig. 4 are interpreted to be due to the forma-tion of co-clusters at about 120–180 °C, whilst the endo-thermic effect is thought to be due to the dissolution of co-clusters with in the latter stages some transformationof co-clusters a phase with long range order, probablythe rod/lath shaped particles observed by Kovarik et al.[23].

DSC curves of the solution treated Al–1.2Mg–1.2Cualloy (alloy 3) at different heating rates are presented in

Fig. 5a. These thermograms show heat effects that haveheat contents well in excess of those observed in alloy 1and 2. Consistent with the well-known theory of thermallyactivated reactions [48], heat effects in these curves shift tohigher temperatures with increasing heating rate. The endtemperature of exothermic effect (T e), which also increaseswith temperature, reflects the stage at which the drivingforce for further reaction has decreased to zero [49]. Thisis the stage at which the average composition of soluteatoms in the Al-rich phase has reached the metastable sol-vus. As T e increases with heating rate, and the solvus con-centrations increase with temperature, the total heat

contents of the exothermic effects decrease with increasingheating rate [50]. On ageing for up to 24 h at room temper-ature, the first exothermic effect in the DSC curves of thesolution treated alloy 3 (Al–1.2Cu–1.2Mg) gradually disap-pears. 3DAP analysis of these alloys presented elsewhere[7,51] shows that whilst immediately after quenching theMg atoms are randomly distributed, Mg atoms becomeclustered, and from analysis of clusters that are larger thanabout 10 atoms, the formation of Cu–Mg co-clusters is

identified. No other phases can be identified, and hencethe low temperature exothermic reaction is attributed tothe formation of Cu–Mg co-clusters. DSC curves of thesolution treated alloy 4 (Al–1.8Cu–1.5Mg) at differentheating rates are presented in Fig. 5b. These thermogramsshow heat effects at similar positions to those in alloy 3,whilst the heat contents of the effects are on average about40% larger. 3DAP analysis similar to that described foralloy 3 indicated that also for alloy 4 the low temperatureexothermic effect was due to the formation of Cu–Mg co-clusters [7].

The measured heat contents of the effects and isothermalageing temperatures for the alloys are presented in Table 2.A range of data on DQ(T e) presented in the literature [52– 54] on a number of alloys is also included in this table.

5.2. Comparing thermodynamic data with the model 

To compare the measured DQ data with the regularsolution model for dimers, model predictions were made

using Eqs. (1)–(3) with Eqs. (7)–(9). In order to achieve

-0.02

0.00

100 150 200 250 300 350 400 450

0min 10min

60min 200min

400min 1000min

2000min 4000min

0 s

600 s

12000 s

24000 s

60000 s120000 s

240000 s

3600 s

Temperature (ºC) -

   H   t  e  a   t   F   l  o  w   (   W   /  g   )  -   >

>

Fig. 4. DSC curves of the solution treated alloy 2B (Al-3.1Mg–0.17Cu)after solution treatment and after ageing for up to 240ks (60.7 h) at

140 °C.

-0.60

-0.40

-0.20

0.00

0.20

0.40

0.60

0.80

Temperature (°C )

   H  e  a   t   F   l  o  w    (

   J   /  g .   °

   C   )

2.5°C/min10°C/min20°C/min40°C/min80°C/min120°C/min150°C/min

-0.60

-0.40

-0.20

0.00

0.20

0.40

0.60

0.80

0 100 200 300 400 500

0 100 200 300 400 500 600

Temperature (°C )

   H  e  a   t   F   l  o  w    (   J

   /  g .   °

   C   )

5°C/min10°C/min20°C/min40°C/min80°C/min100°C/min120°C/min

a

b

Fig. 5. DSC curves of a) the solution treated alloy 3 (Al–1.2Cu–1.2Mg) at(from left to right) heating rates 2.5, 10, 20, 40, 80, 120 to150 K/min; andb) the solution treated alloy 4 (Al–1.8Cu–1.5Mg) at (from left to right)heating rates 5, 10, 20, 40, 80, 100 and 120 K/min.

2382 M.J. Starink, S.C. Wang / Acta Materialia 57 (2009) 2376–2389

Page 8: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 8/14

these model predictions we first needed to determineDH A–B. From each data line in Table 2 a value of DH A–B

can be obtained, we take DH A–B as the average from theexperiments that are expected to yield the highest accuracy:DSC experiments at 5–40 °C/min for the alloy with thehighest xCu + xMg. This provides DH A–B = 34.5 ± 0.5 kJ/

mol. The predicted DQ values are presented in Table 2and graph of predicted vs measured values of DQ consider-ing all reactions which completed within 20 min forT e > 100 °C and completed within 7 days for T e 6 60 °Cis provided in Fig. 6. Plots of DQ vs T e for alloys 3 and 4are provided in Fig. 7. The correspondence between theregular solution model for dimers and the thermodynamicdata obtained from calorimetry is excellent for all alloys.

The validity of the thermodynamic model was furtherchecked by comparing predictions for DQ using the moregeneral equation Eq. (5) and optimising a, D H MmAaBb andC  (b is fixed at 1, this is not a limitation as there are onlythree independent parameters). Best fits were obtained witha close to 1, which confirms that clusters have Cu:Mg ratiosclose to 1. Solvus predictions with optimised a, DH c and C 

for the alloys were within about 2 °C of the one obtainedwith the regular solution model for dimers.

Additional DSC experiments were performed on twoalloys that are right at the limit of stability of co-clustersat typical formation temperatures: Al–0.08Cu–1Mg (at%)and Al–0.08Cu–2Mg (at%). For these alloys predicted heatevolution are 0 and 0.4 J/g, respectively. Observed heat of dissolution of co-clusters are right at the detection limit,at 0.2 ± 0.2 J/g averaged for the two alloys. These observa-tions should be compared with the data for alloy 2b, for

which co-cluster formation and dissolution was readily

observed. Thus, these observations provide further supportfor the model predictions on stability limits of co-clustersfor these alloys with very low Cu content.

Considering the typical measurement error, the level of agreement between calorimetry data and predictions of the thermodynamics based on a regular solution model inFigs. 6 and 7 is excellent. These figures contain thermody-namic data for a wide range of alloys for the full range of temperatures at which co-cluster formation is normally

observed (20–200 °C). To prove the regular solution model

Table 2Magnitude of the heat contents of exothermic precipitation effect and heat contents of endothermic dissolution effect ( DQ) with gross compositions ( xgCu, x

gMg, xgMn). In the first column ‘DSC exo’ stands for an exothermic effect observed in a linear heating experiment, ‘DSC endo’ stands for an endothermic

effect observed in a linear heating experiment, IC stands for an exothermic effect observed in an isothermal calorimetry experiment. T e stands for the endtemperature of the exothermic effect in a ‘DSC exo’ experiment, or the ageing temperature in a ‘DSC endo’ experiment, or the isothermal ageingtemperature during the IC experiment.

Experiment Reference xgCu x

gMg x

gMn T e (°C) DQ (J/g) DQ (J/g)

Measured PredictedIC Zahra et al. [52] 0.0025 0.047 0.0007 60 3.5 3.15DSC exo Verlinden and Zahra [53] 0.0025 0.047 0.0007 195 2.9 2.72IC Zahra et al. [52] 0.0023 0.033 0.0004 60 3.5 2.89IC Zahra et al. [52] 0.0040 0.033 0.0008 60 6 5.04DSC exo Verlinden and Zahra[53] 0.0040 0.033 0.0008 180 4.4 4.56IC Zahra et al. [52] 0.0090 0.014 0.0000 30 11.7 11.4IC Zahra et al. [52] 0.0110 0.017 0.0000 30 13.5 13.9DSC exo This work 0.0007 0.013 0.0000 170 0.2 0DSC exo This work 0.0007 0.013 0.0000 170 0 0IC Zahra et al. [52] 0.0025 0.047 0.0007 0 0 3.16IC Verlinden and Zahra [53] 0.0040 0.033 0.0008 0 0 5.06DSC exo Charai et al. [54] 0.0087 0.014 0.0000 150 9.4 9.81DSC exo Charai et al. [54] 0.0090 0.015 0.0000 165 9.3 9.77DSC exo This work 0.0121 0.012 0.0020 111 14.1 12.5

DSC endo This work 0.0018 0.030 0.0000 180 2 1.74DSC This work 0.0018 0.030 0.0000 215 0 0.00DSC endo This work 0.0018 0.030 0.0000 180 0.5 0.00DSC endo This work 0.0017 0.032 0.0012 140 2.0 1.95DSC endo This work 0.0017 0.032 0.0012 200 0.7 1.45

0

5

10

15

20

0 2 4 6 8 10 12 14 16 18 20

Measured (J/g)

   P  r  e   d   i  c   t  e   d   (   J   /  g   )

DSC and IC for Cu<0.9at%

Series3

DSC for Cu>0.9at%

Fig. 6. Measured vs. predicted DQ from calorimetry experiments for a

range of Al–Cu–Mg alloys. a = 1, DH c = 34.5 kJ/mol. Â, DSC data onalloys 3 and 4; h, all other data.

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2383

Page 9: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 9/14

is not just a fortuitous fit but physically realistic, we willshow in Section 6 that the present interpretation is consis-tent with published microstructural data.

5.3. Strength and hardness data

The yield strength of an alloy that contains co-clusters isdominated by the contribution due to clusters, with furtherminor contributions due to grain boundaries, dislocationsand solute in the Al-rich phase. A general approximationof the superposition of strengthening mechanisms is [55]:

r y  ¼ r gb þ M ðs sol þ sd þ sclÞ¼ r gb þ M ðs sol þ sd þ ðDs sro þ Dsm þ DschÞÞ ð19Þ

where rgb is the contribution due to grain boundaries sd isdue to stored dislocations (introduced by working) and ssolis due to dissolved elements. The method for obtaining rgbis outlined in [55], it provides rgb values that are less than1 MPa for the relatively coarse grained alloys consideredhere. In the alloys considered, sd is zero for solution treatedand aged alloys and for T351 (solution treated stretchedand aged alloys) it is obtained from the model outlined in[56], which provides sd values of about 5 MPa. ssol is ob-tained form the model outlined in [57]. To verify the modelwe will compare yield strength predictions with data for arange of solution treated and subsequently aged Al–Cu– Mg alloys obtained from this work and from [58–66].(Strength of rolled Al–Cu–Mg alloys is anisotropic. Whereavailable we will use the average of L and LT tests; if onlyL is available we will consider that the average of L and LTyield strengths, ry,L and ry,LT, is about 0.95ry,LT [67].) Theresults, presented in Table 3, show a very good correspon-dence: the typical accuracy (root mean square error) is11 MPa for data ranging from 110 to 350 MPa. This limitedlevel of deviations is thought to be mostly due to smalluncertainties not related to the present strength model thatinclude uncertainties in alloy composition (a 0.1wt% change

in Cu content causes 6 MPa change in predicted ry), limited

measurements inaccuracies and the anisotropy of yieldstrength introduced by texture of rolled alloys.

6. Discussion

Comparison of the thermodynamic model for dimers(regular solution model) and the strength model with calo-rimetry and yield strength data of well over a dozen Al– Cu–Mg alloys spreading over a wide range of compositionsshows an excellent correspondence, with any remaining

deviations likely to be mainly (or exclusively) caused bylimited experimental errors. This has been achievedthrough determining just one parameter DH A–B. For otheralloy systems that may contain co-cluster no consistent setof calorimetry and strength data could be found and hence,at present, we cannot test the models for other alloys. How-ever it appears clear from the present model that for arange of ternary or higher order alloys that are processedthough a supersaturated solid solution involving two ormore alloying elements decomposition will progressthrough a co-cluster stage. In the following we will focusour discussion on co-clusters in the Al–Cu–Mg system.

In the following discussion we will first discuss possiblevariants of the thermodynamic model, cluster size and clus-ter composition. Secondly, we will aim to show that allpublished microstructural data and quantum mechanicmodelling is consistent with the present models. Thirdlywe will consider possible other strengthening models.And finally we will show how the models can quantitativelyexplain the phenomenon of secondary ageing in Al–Li–Cu– Mg alloys.

6.1. Cluster composition and further evolution

As the fitting of the broader thermodynamic model

(Eq. (5)) provides a:b = 1, the basic unit element of the

0

2

4

6

8

10

12

14

16

18

0 100 200 300 400

T (ºC)

   Q   (   J   /  g   )

      Δ

Series1

Series2

Series3

Series4

Series5

Fig. 7. Measured DQ (symbols) and predicted DQ (lines) as a function of end temperature for alloys 3 and 4. a = 1, DH c = 34.5 kJ/mol.

Table 3Predicted and measured yield strengths of various solution treated andsubsequently aged Al–Cu–Mg alloys. Ageing is conducted to a state whereco-cluster formation is completed: 4–10 days at 25 °C and about 5 min at150 and 180 °C.

 xgCu (at) xgMg (at) ry measured

(MPa)ry modelled(MPa)

T  (°C) Ref.

0.0179 0.0139 315 295 25 This work0.0119 0.0117 265 273 25 This work0.0106 0.0160 250 262 30 Wil [58]0.0106 0.0128 255 258 20 Pol [59]0.0050 0.0170 150 148 150 Rav [60]0.0080 0.0170 200 204 150 Rav [60]0.0110 0.0170 245 246 180 Rav [60]0.0187* 0.0160* 325 343* 25 ASM [62]0.0183 0.0148 348 346 25 Sri [63]0.0179 0.0145 328 316 25 This work0.0025 0.0450 130 140 180 Rat [64]0.0024 0.0320 95 114 180 Cou [65]0.0022 0.0427 112 126 180 Cou [65]0.0169 0.0108 277 289 25 Fin [66]

*

The median 2024 composition with 0.1 wt% Fe is assumed.

2384 M.J. Starink, S.C. Wang / Acta Materialia 57 (2009) 2376–2389

Page 10: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 10/14

co-cluster is Cu1Mg1Alp or multiples of these. In terms of the regular solution model this leaves open various possi-bilities, for instance complexes of 2 Cu atoms with 2 Mgatoms, with the Cu and Mg atoms not situated as nearestneighbours but separated by a certain number of Al atoms.Examples are given in Fig. 1. However, the excellent corre-

spondence of the thermodynamic model for dimers and thestrengthening model for dimers indicates that the mainreaction involves the formation of the Cu–Mg dimers:one Cu and one Mg atom in nearest neighbour positions.However this in itself does not exclude the possibility of larger co-clusters; and in fact all co-cluster constellationswhere DH  for co-cluster formation is essentially relatedto a chain of bonds such as illustrated in Fig. 1 is consistentwith the measured enthalpy changes and strength data. Iso-thermal calorimetry data shows that exothermic heat evo-lution during room temperature hardening peaks duringthe hardening, and reduces to zero by the time the hardnessstabilises after about 48 h at room temperature. The latter,

whilst consistent with the present models for dimers as wellas extended chains where DH  for co-cluster formation isessentially related to a chain of bonds, also indicates thatany further growth of co-clusters does not have a signifi-cant impact on the enthalpy of the alloy. Hence, the occa-sional occurrence of larger co-clusters (larger than 10atoms, in some cases larger than 100 atoms) should be con-sidered as a coagulation of smaller co-clusters; with thecoagulation not influencing the strength or DH  to a signif-icant extend.

In these coagulated dimers some relaxation of theatomic positions will occur and the model for a needle/lath

like phase appearing after 24–96 h at 180 °C [23] can in factbe considered as a coagulation of dimers like in Fig. 1c withsome relaxation of atoms positions.

6.2. Nanostructure data and quantum mechanical modelling 

Whilst atom probe analysis cannot show the existence of individual dimers, data is fully consistent with the abovemodel. In particular, APFIM on Al–Cu–Mg based alloysshow that [7]:

j During ageing after solution treatment co-clusters areformed.

j Clusters of sizes that can be reliably detected (typicallymore than $10 atoms) do occur but only a small frac-tion of the Cu and Mg atoms are associated with theselarger clusters.

j The larger clusters have Cu:Mg ratios that vary, witha median value of about 1.

Thus atom probe data is fully consistent with the presentmodels on dimer formation, and the experiments showexactly the features that can be expected. Also NuclearMagnetic Resonance (NMR) [10] and positron annihila-tion lifetime spectroscopy (PALS) [13] are sensitive to

Cu–Mg dimer formation. NMR and PALS studies on

Al–Cu–Mg alloys have indeed shown that after the rapidhardening reaction the chemical environment of Cu atomshas changed [10,13], but the published analysis of theexperiments fails to identify what this change entails.Microstructural data for two broad groups of experiments(near room temperature and near 150 °C) are presented in

Table 4. The table includes TEM and selected area diffrac-tion (SAD) on an Al–4wt%Cu–1.5wt%Mg alloy that wassolution treated, quenched and aged at room temperaturefor about a year. This confirmed that no additional diffrac-tion phenomena are observed in SAD and no precipitatescan be observed in bright field TEM imaging.

The data in Table 4 is fully consistent with the presentmodels for co-clusters, with the exception of one observa-tion by Reich et al. [68] who failed to detect Cu–Mg co-clusters in a sample aged for 1 min at 150 °C (by they dodetect them after 5 min). In the absence of any data sup-porting this one finding we consider the weight of evidenceclearly supports the present models, and we will disregard

this one finding. We further investigated literature onTEM, HREM, SAD in TEM, APFIM, 3DAP, calorimetry,NMR and PALS experiments for ageing times withinabout 10 times the time to reach the plateau hardness (ortime to stabilise the enthalpy) in about 50 papers (encom-passing [7,12,19,20,23] and references therein, and refer-ences [10,11,13,15–22,44,51–53,64]), and concluded thatthe present model is consistent with all the available data.In particular it should be noted that no microstructureinvestigation ever showed evidence for precipitates (otherthan co-clusters suggested by APFIM, 3DAP) for ageingtimes within about 10 times the time to reach the plateau

hardness (or time to stabilise the enthalpy).This assessment further indicates that the term GPB

zones, associated with early decomposition and hardeningin Al–Cu–Mg alloys in most of the pre 1995 papers citedin this work, does not reflect accurately the processesinvolved, and is ultimately obfuscating or, depending onwhich of the definition of GPB zones is used, evenmisleading.

Next the first principles model prediction (VSP) wasassessed. A first principles total energy calculation (FP-CVM) of 4 atom tetrahedrons provides an enthalpy of mix-ing of 30 kJ per mole Cu–Mg dimers [25]. This is in goodagreement with the present model and analysis, and thesmall deviation can be due the limitations in accuracy pro-vided by this 4 atom tetrahedrons calculation. We may alsocompare our results with the enthalpy of mixing for thecoherent rod/lath-like structures formed in a later stageof the ageing (24–96 h at 180 °C) in an Al–0.4at%Cu– 3at%Mg alloy [23]. For the coherent rod/lath-like struc-tures identified by Kovarik et al. [23], 5 Â 5 Â 1 and6 Â 6 Â 1 supercell calculations determine an enthalpy of formation of 48 kJ per mol solute atoms in the GPB struc-ture [23]. Thus these coherent rod/lath-like structures aremore stable then the Cu–Mg dimers and should be consid-ered a stage of the ageing of the alloys that is subsequent to

the co-cluster formation.

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2385

Page 11: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 11/14

An alternative treatment for thermodynamics of clusterssuggested in the literature [24] involves a quasi-chemicalmodel which takes nearest neighbour interactions intoaccount. Analysis of this model showed that it cannotexplain all the experimental observations. Particularly, thismodel cannot reconcile the observations that immediatelyafter water quenching no Cu–Mg co-clustering is detected,with the enthalpy of formation of Cu–Mg co-clusteringbeing of the order determined here.

6.3. Strength models for co-clusters: comparison with

 previous models

The present model for co-cluster strengthening indicatesthat the yield strength of alloys which contain only co-clus-

ters as strengthening precipitates is in good approximation

proportional to the amount of the co-clusters. This is differ-ent in previous models for co-cluster strengthening (and thevast majority of other precipitation strengthening mecha-nisms) which predict a square root of the amount (orsquare root of volume fraction) dependency. This allowsus to test the model in the following fashion. We have per-formed hardness tests on an Al–1.2at%Cu–1.2at%Mg alloyduring room temperature ageing for time tNA, and alsoperformed DSC experiments on samples of the same alloysubjected to the same natural ageing treatments. The DSCthermograms are presented in Fig. 8. The total change inheat released during the DSC experiment for each naturallyaged sample, DQ(tNA), was measured and the changeduring ageing, DQ(tNA) À DQ(tNA = 0), was determined.The latter difference, DQ(tNA)

ÀDQ(tNA = 0), should be

proportional to the amount of co-clusters formed during

Table 4Microstructural data on the two broad groups of experiments: ageing experiments performed at room temperature and near 150 °C, for times up to about10 times the time required to reach the hardness plateau.

Alloy (at%) T  (°C) t Experiment Observation

This work Al–1.9Cu–1.6Mg 25 10000 h TEM No precipitates resolvableThis work Al–1.9Cu–1.6Mg 25 10000 h SAD No diffraction effects due to precipitates

Starink et al.[7]

Al–1.9Cu–1.6Mg and Al– 1.2Cu–1.2Mg

25 0–200 h Calorimetry H  decreases strongly*

, H  stabilises at about 2 h

Starink et al.[7]

Al–1.9Cu–1.6Mg and Al– 1.2Cu–1.2Mg

50 0–200 h Calorimetry H  decreases strongly*, H  stabilises at about 15 h

Fink et al.[66]

Al–1.7Cu–(0.5–1.5)Mg 30 0–10 h Dilatometry Expansion

Fink et al.[66]

Al–1.7Cu–(0.5–1.5)Mg 31 0–10 h Dilatometry Increased resistivity

Staab et al.[12]

Al–2.0Cu–1.8Mg–0.6Mn 25 120 h XANES Consistent with Cu–Mg bonds, not with Cu in solution,or Cu–Cu clustering;

Starink et al.[7]

Al–1.2Cu–1.2Mg 25 0.5–3.5 h 3DAP-statistical

Onset of co-clustering happens between 1 and 2 h

Starink et al.[7]

Al–1.2Cu–1.2Mg 25 2–30 h 3DAP-imaging

Clusters with more than 7 atoms are detectable, butnumbers insignificant

Bastow [10] Al–1.05Cu–1.7Mg 21 0–645 h NMR Change chemical environment of atoms from

$10 min

to $30 hRinger et al.

[21]Al–1.1Cu–1.7Mg 20 24 h Hardness Hardening to plateau completed at 24 h

Starink et al.[7]

Al–1.9Cu–1.6Mg and Al– 1.2Cu–1.2Mg

25 0–200 h Hardness Hardening to plateau occurs between 0.01 and 2 h

Reich et al.[68]

Al–1.7Mg–1.1Cu 150 1 min TEM No precipitates resolvable at 1 min

Reich et al.[68]

Al–1.7Mg–1.1Cu 150 1 min SAD No precipitates at 1 min

Reich et al.[68]

Al–1.7Mg–1.1Cu 150 1 min Hardness Hardness reaches plateau within 1 min

Reich et al.[68]

Al–1.7Mg–1.1Cu 150 1 min 3DAP Cu:Mg ratio in larger clusters about 3:5

Reich et al.[68]

Al–1.7Mg–1.1Cu 150 1–60 min 3DAP Cu–Mg co-clusters detected at 5 min, not detected at1 min

Ringer et al.[20]

Al–1.7Mg–1.1Cu 150 5 min AP Cu–Mg co-clusters detected at 60 min

Ringer et al.[20]

Al–1.7Mg–1.1Cu 150 5 min TEM No precipitates resolvable at 5 min

Ringer et al.[21]

Al–1.1Cu–1.7Mg 130 linear heating$10 min

Calorimetry H  decreases strongly*

This work Al–1.2Cu–1.2Mg 120 linear heating$10 min

Calorimetry H  decreases strongly*, H  stabilises at 120 °C duringheating at 10 K/min

* Change in H  is about half of that due to formation of the equilibrium phase, the S phase.

2386 M.J. Starink, S.C. Wang / Acta Materialia 57 (2009) 2376–2389

Page 12: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 12/14

ageing, and the amount of co-clusters thus obtained is plot-ted as a function of the hardness in Fig. 9. (One furtherdata point for Al–1.9at%Cu–1.6at%Mg aged for 100 h isadded.) The graph of  f  vs HV shows a linear correlation,which further supports the present strengthening model,whilst clearly showing strength increments are not propor-tional to f 1/2.

In several works (e.g. [38]), including some from thepresent group (e.g. [7,57]), strengthening by co-clustershas been modelled by considering only modulus hardeningis effective. Eq. (14) (the simplified form for modulus hard-ening) was applied in these works, and through fitting of lclgenerally good results were obtained. Whilst the presentwork indicates that this approach is mistaken in terms of the identification of the fundamental processes at work, itcan be seen that fitting a value for lcl higher than the actualvalue of lcl can provide an effective fit for a limited range of 

data.

We further note that the present co-cluster hardeningmodel provides an explanation for a range of observationsin ternary and higher order alloys. For example, throughcombining Eqs. (6), (9), (13), and (19), and assuming solu-tion strengthening is substantially less than co-clusterstrengthening, the increase in strength of an alloy on for-

mation of co-clusters is proportional to the change inenthalpy during co-cluster formation. Such a proportional-ity is found for Al–1.7%Li–Cu–Mg alloys that wereexposed for 1000 h at 70 °C following an solution treatingand an ageing treatment at 150 °C [69] and the proportion-ality constant is consistent with the present model. Thissecondary ageing treatment does not cause any changesin microstructure that can be detected by TEM. This showsthat the magnitude of secondary age hardening in this alloycan be quantitatively explained by the formation of co-clusters.

It is also noted that the present models show that co-clusters impart both a significant strength and, as a result

of the strong decrease resistance to dislocation movementon passage of a one dislocation, they also cause an unusu-ally high propensity for shear localisation. In fatigue load-ing conditions the shear localisation would cause apropensity for roughness induced crack closure (RICC)[8], provided grains are reasonably large. Thus an alloywith high amounts of co-clusters would be ideal for appli-cations requiring fatigue resistance and strength, and opti-misation of these mechanical properties would be achievedthrough maximising the co-cluster content. Publishedcalorimetry and APFIM work on a range of alloys haveindicated that of the light fcc metallic alloys Al–Cu–Mg

have the highest co-cluster content. From this informationwe can now determine the composition of the optimumalloy for specific strength and fatigue resistance in the fol-lowing way:

 – Our model shows that underaged Al–Cu–Mg alloyswith Cu:Mg content close to 1 and Cu and Mg solutelevels close to the maximum solubility should providemaximum co-cluster content. Using phase diagramsof ternary alloys (see e.g. [27]) this indicates an Al– 1.6Cu–1.6Mg alloy.

 – We also consider that Mn additions at levels of about0.2 at% are effective in promoting recrystallisationthrough formation of T-Al20Cu2Mn3 dispersoids[71], which are to a large extend not dissolvable inthe Al-rich phase. Thus Mn additions will removesome Cu from solid solution and also Fe impuritieswill remove some Cu from solid solution (through for-mation of x-Al7Cu2Fe); and it is clear that the bestalloys for specific strength and fatigue resistance(and commercially viable impurity levels) should beAl–1.8Cu–1.6Mg–0.2Mn–0.05Fe alloys in underagedconditions.

The latter compositions and heat treatment are essen-

tially identical to 2x24-T3 alloys, which are dominant in

-0.04

-0.02

0

0.02

0.04

0.06

0 50 100 150 200 250

Temperature (°C)

   H  e  a   t   f   l  o  w   (  w   /  g   )

0.08h1 h4 h10 h170 h

Fig. 8. DSC thermograms for quenched and subsequently room temper-ature aged Al–1.2at%Cu–1.2at%Mg alloy. Depicted are thermogramsafter ageing times of (in order of decreasing exothermic heat effect) 0.08, 1,4, 10 and 170 h.

40

60

80

100

120

140

160

0 0.005 0.01 0.015 0.02 0.025

   H   V

Fraction of Cu atoms

 Al-1.2Cu-1.2Mg

 Al-1.9Cu-1.6Mg

Fig. 9. Measured Vickers hardness vs the fraction of Cu and Mg atoms inco-clusters (for DH c = 34.5 kJ/mol).

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2387

Page 13: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 13/14

applications requiring a high specific strength and fatigueresistance, such as airplane fuselage and lower wing. Thusthis analysis provides a rationale for alloy optimisation,linking atomic scale processes to properties at the compo-nent scale (wings and fuselage).

7. Conclusions

A thermodynamic model for co-clusters based on a sin-gle interaction energy of dissimilar nearest neighbour inter-action energy is presented. Also a model for thestrengthening due to these co-cluster dimers is derived.The model includes a new treatment of (short-) orderstrengthening relevant to these co-clusters and furtherencompasses modulus hardening and chemical hardening.Evaluation shows that in general (short-) order strengthen-ing will be the main strengthening mechanism. The modelis tested against data on Al–Cu–Mg alloys. It is shownthat:

 – The thermodynamic model can fully explain pub-lished and new data on the heat evolution due to co-cluster formation. The Cu–Mg interaction enthalpyDH A–B is determined as 34.5 ± 0.5 kJ/mol.

 – The models can quantitatively predict the very firststage of hardening and associated heat evolution inAl–Cu–Mg alloys with Mg:Cu ratio between 20 and0.5, and Cu contents between about 0.1 and 2 at%.The process of co-cluster formation is consistent withpublished data on microstructural evolution for rele-vant Al–Cu–Mg alloys.

 – The model is successfully tested against data onstrength and hardness of a wide range of Al–Cu–Mgalloys treated at temperatures between 20 and 220 °C.

 – The term GPB zones, often associated with earlydecomposition and hardening in Al–Cu–Mg alloys,does not reflect accurately the processes involved,and is ultimately obfuscating or, depending on whichof the definition of GPB zones is used, evenmisleading.

 – Co-cluster formation involving dimers causes strengthincrements that are in good approximation propor-tional to the change in enthalpy of the alloy, and theproportionality constant is derived. This proportion-ally is seen in Al–Li–Mg–Cu alloys exposed at70 °C, which strongly indicates that strengtheningduring low temperature exposure is due to formationof 2-atom co-clusters.

Acknowledgements

The following people are gratefully acknowledged: Ms.A. Dion, and Drs. J.L. Yan, Z. Zhu and N. Gao for per-forming selected DSC experiments, Dr. N. Gao for per-forming tensile tests on a 2024 alloy, Dr. G. Mahon for

thermomechanical processing of Alloy 2b, Drs. G. Mahon,

S. Court and L. Kovarik for discussions on low Cu Al– Mg–Cu–Mn alloys. QinetiQ (Farnborough, UK), Alcan(former Banbury Labs), Airbus UK and Prof M.J. Millsand Dr. L. Kovarik are gratefully acknowledged for pro-viding alloys.

Appendix A. Undissolved intermetallics

Insoluble or undissolved particles and dispersoids, suchas x-Al7Cu2Fe, S-Al2CuMg, T-Al20Cu2Mn3, Al12(Fe,Mn)3Si and Al6(Fe,Mn), which form during solidificationor homogenisation, can be present in Al–Cu–Mg typealloys with Mn, Fe and Si additions and impurities[26,27]. Since the presence of x, S and T phases will removesome Cu and Mg from the solid solution, it is necessary tocalculate the effective solute concentration in the matrixafter solution heat treatment.

The atomic fraction of undissolved x-Al7Cu2Fe can becalculated based on the solubility of Fe in Al matrix, sincethe solubility of Fe is not significantly influenced by theadditions of Cu or Mg [70]. At typical T SHT for 2024 alloys,which is about 500 °C, the maximum solubility of Fe in Al, xeFe, is 0.0055 wt.% [71]. The amount of x-Al7Cu2Fe in thealloy is then given by:

 xx ¼ 10ð xgFe À xeFeÞ ð20Þwhere xgFe is the gross Fe content of the alloy. A similartreatment is applied to obtain the amount of Al20Cu2Mn3.Based on the solvus of Al20Cu2Mn3 in the Al–Cu–Mn sys-tem, at 500 °C the solubility of Mn in Al with a Cu contentof 4 wt%, is about 0.2wt% [70]. Thus the amount of T-

Al20Cu2Mn3 in the alloy is given by: xT ¼ 25=3ð xgMn À xeMnÞ ð21Þwhere xgMn is the gross Mn content of the alloy.

In alloys for which the Cu and Mg contents are close tothe solvus of S phase at the homogenising treatment, someundissolved S phase can be present. FEG-SEM investiga-tion of alloys 2b, 3 and 4 showed that only alloy 4 con-tained undissolved S phase. DSC experiments showedthat this alloy has an S phase solvus temperature withinabout 10 °C of the solution heat treatment temperature,T SHT, whilst for the other two the difference is at least30 °C. Thus we will consider that for all alloys with an Sphase solvus temperature more than 30 °C below T SHT

all S phase is dissolved. For all other alloys we will estimateS phase content from available experimental data. Anaccurate calculation of the amount of S phase in alloy 4is provided by the difference between the amounts of inter-metallic phases detected in alloys 3 and 4. This providesV S = 0.009 for alloy 4. For the 2524 alloy in [59] we canestimate the amount of S phase from optical micrographsin [59] as about half that of our alloy 4.

The effective solute concentration in the matrix aftersolution treatment is obtained by from the difference of the gross content and the amounts in the undissolved

phases.

2388 M.J. Starink, S.C. Wang / Acta Materialia 57 (2009) 2376–2389

Page 14: The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–Mg alloys

7/31/2019 The thermodynamics of and strengthening due to co-clusters General theory and application to the case of Al–Cu–…

http://slidepdf.com/reader/full/the-thermodynamics-of-and-strengthening-due-to-co-clusters-general-theory-and 14/14

References

[1] Pereloma EV, Shekhter A, Miller MK, Ringer SP. Acta Mater2004;52:5589.

[2] Pereloma EV, Timokhina IB, Russell KF, Miller MK. Scripta Mater2006;54:471.

[3] Moody MP, Stephenson LT, Liddicoat PV, Ringer SP. Microsc Res

Techn 2007;70:258.[4] Stephenson LT, Moody MP, Liddicoat PV, Ringer SP. MicroscMicroanal 2007;13:448.

[5] Honma T, Yanagita S, Hono K, Nagai Y, Hasegawa M. Acta Mater2004;52:1997.

[6] Serizawa A, Hirosawa S, Sato T. Metall Mater Trans A2008;3:243–51.

[7] Starink MJ, Gao N, Davin L, Yan J, Cerezo A. Phil Mag2005;85:1395.

[8] Kamp N, Gao N, Starink MJ, Sinclair I. Int J Fatigue2007;29:869–78.

[9] Williams JC, Starke EA. Acta Mater 2003;51:5775–99.[10] Bastow TJ. Phil Mag 2005;85:1053.[11] Bastow TJ, Hill AJ. Mater Sci For 2006;519–521:1355–60.[12] Staab TEM, Haaks M, Modrow H. Appl Surf Sci 2008;255:132–5.

[13] Klobes B, Staab TEM, Haaks M, Maier K, Wieler I. Phys Stat Sol2008;2:A61–6.

[14] Hardy HK. J Inst Metals 1954–55;83:17.[15] Vietz JT, Polmear IJ. J Inst Metals 1966;94:410.[16] Takahashi T, Sato T. J Jpn Inst Light Metals 1985;35:41.[17] Nagai Y, Murayama M, Tang Z, Nonaka T, Hono K, Hasegawa M.

Acta Mater 2001;49:913.[18] Abis S, Massazza M, Mengucci P, Riontino G. Scr Mater

2001;45:685.[19] Starink MJ, Cerezo A, Yan J, Gao N. Philosoph Mag Lett

2006;86:243.[20] Ringer SP, Sakurai T, Polmear IJ. Acta Mater 1997;45:3731.[21] Ringer SP, Caraher SK, Polmear IJ. Scr Mater 1998;39:1559.[22] Zahra A-M, Zahra CY. Phil Mag Lett 2002;82:9.[23] Kovarik L, Court SA, Fraser HL, Mills MJ. Acta Mater

2008;56:4804–15.[24] Zhu Aiwu, Gable BM, Shiflet GJ, Starke EA. Acta Mater

2004;52:3671.[25] Zhu Jr Aiwu, Starke EA, Shiflet GJ. Scripta Mater 2005;53:35.[26] Wang SC, Starink MJ. Int Mater Rev 2005;50:193.[27] Bron F, Besson J, Pineau A. Mater Sci Eng A 2004;380:356.[28] Ardell AJ. Metall Trans A 1985;16:2131.[29] Schlieser C, Nembach E. Acta Metall Mater 1995;43:3983.[30] Gerold V, Gudladt HJ, Lendvai J. Phys Stat Sol A 1992;131:509.[31] Nembach E. Scripta Mater 1997;36:1409.[32] Jeon SM, Park JK. Acta Mater 1996;4:1449.[33] Buchner AR, Pitsch W. Z Metallkde 1985;76:651.[34] Flinn PA. Acta Metall 1958;6:631.[35] Kostorz G, Schonfeld B. Chimia 2001;55:517.

[36] Mohri Tetsuo, Miyamoto Hirotsugu, Terada Yoshihiro, SuzukiTomoo. Mater Sci Eng A 1999;267:151.

[37] Cartaud L, Guillot J, Grilhe J. Proceedings of ICSMA IV, Nancy, 30August–3 September 1976, vol. 1, (Ed. Laboratoire de Physique duSolide ENSMIN-INPL, Nancy) 1976. p. 214.

[38] Gomiero P, Brechet Y, Louchet F, Tourabi A, Wack B. Acta MetallMater 1992;40:857.

[39] Smithells CJ. Metals reference book. 7th ed. London: Butterworths-Heinemann; 1992.

[40] Nembach E. Phys Stat Sol A 1983;78:571.[41] Wang SC, Zhu Z, Starink MJ. J Microsc 2005;217:174.[42] Zhu Z, Starink MJ. Mater Sci Eng A 2008;488:125.[43] Starink MJ. Int Mater Rev 2004;49:191.[44] Starink MJ, Dion A. Thermochim Acta 2004;417:5.[45] Yan J. Ph.D. Thesis, University of Southampton; 2006.[46] Kovarik L, Gouma PI, Kisielowski C, Court SA, Mills MJ. Mater Sci

Eng 2004;A387–A389:326.[47] Kovarik L, Gouma PI, Kisielowski C, Court SA, Mills MJ. Acta

Mater 2004;52:2509.[48] Starink MJ. Thermochim Acta 2003;404:163.[49] Khan IN, Starink MJ. Mater Sci Techn 2008;24:1403–10.[50] Starink MJ, Gregson PJ. Scripta Metall Mater 1995;33:893.[51] Davin L. Ph.D. Thesis, Oxford University; 2004.[52] Zahra A-M, Zahra C, Verlinden B. Phil Mag Lett 2006;86:235.[53] Verlinden B, Zahra A-M. Mater Sci Forum 2003;426–432:423.[54] Charai A, Walther T, Alfonso C, Zahra AM, Zahra CY. Acta Mater

2000;48:2751.[55] Starink MJ, Deschamps A, Wang SC. Scripta Mater 2008;58:377.[56] Khan IN, Starink MJ, Yan JL. Mater Sci Eng A 2008;472:66.[57] Zhu Z, Starink MJ. Mater Sci Eng A 2008;489:138.[58] Wilson RN, Moore DM, Forsyth PJE. J Inst Met 1967;95:177.[59] Polmear IJ. Trans Met Soc AIME 1964;230:1331.[60] Raviprasad K, ICAA9, 2–5 August 2004, Brisbane, Australia [poster].[61] Winkelman GB, Raviprasad K, Muddle BC. Mater Sci Forum

2002;396:1037.[62] Davis JR, editor. ASM specialty handbook aluminum and aluminum

alloys. ASM International; 1993.

[63] Srivatsan TS, Kolar D, Magnusen P. Mater Des 2002;23:129.[64] Ratchev P, Verlinden B, De Smet P, Van Houtte P. Acta Mater1998;46:3523.

[65] Court SA, Kovarik L, Mills MJ. Unpublished research; 2006.[66] Fink WL, Smith DW, Willey LA. In: ‘Age Hardening of Metals’,

Symposium on precipitation hardening, Oct 23–27, 1939. Chicago:American Society for Metals; 1940.

[67] Weilong Hu. Int J Plast 2007;23:620.[68] Reich L, Ringer SP, Hono K. Phil Mag Lett 1999;79:639.[69] Katsikis S, Noble B, Harris SJ. Mater Sci Eng 2008;A485:613.[70] Villars P, Prince A, Okamoto H, editors. Handbook of ternary alloy

phase diagrams. Materials Park, OH: ASM International; 1994.[71] Phillips HWL. The constitution of alloys of aluminium, copper and

iron. J Inst Met 1953–54;82:197.

M.J. Starink, S.C. Wang/ Acta Materialia 57 (2009) 2376–2389 2389