the interaction of grain refinement and ageing in ... · the interaction of grain refinement and...

10
The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson , C. Paa-Rai Manchester Materials Science Centre, Grosvenor Street, Manchester M1 7HS, UK article info Article history: Received 16 April 2015 Revised 8 May 2015 Accepted 9 May 2015 Available online 31 May 2015 Keywords: Precipitation Strengthening Magnesium alloys Modelling abstract Magnesium–zinc–zirconium (ZK) alloys are precipitation strengthened by the addition of zinc and grain refined by zirconium. In ZK alloys these mechanisms interact since some Zn can be precipitated into Zn–Zr intermetallics. This reduces the fraction of age hardening precipitates. An experimental study has confirmed the presence of these two particle types in ZK60, a commercial Mg–Zn–Zr alloy. A model has been developed to consider precipitation and strengthening in ZK alloys accounting for this interac- tion. It has been shown that whilst Zr additions will increase strength by grain refinement this is balanced by a loss of age hardening response. It is predicted that if the same fine grain size could be achieved without Zr then a significant strength increase of around 15% would be expected. Therefore, in wrought applications where the final grain size is determined by recrystallisation, the use of Zr grain refiner in ZK alloys does not lead to optimum strengthening. The model was then applied to explore strategies to increase strengthening in Mg–Zn alloys. Ó 2015 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). 1. Introduction Grain refinement and age hardening (precipitation strengthen- ing) are two metallurgical phenomena that are exploited in magne- sium alloys to achieve the desired microstructure and mechanical properties. The refinement of grains in the as-cast condition gener- ally leads to an improvement in mechanical properties and struc- tural uniformity in most metals and alloys. The use of a grain refining addition to promote the formation of a fine grain structure is particularly important for magnesium alloys produced under conditions where the solidification rate is not very rapid, such as encountered in sand casting, permanent mould casting, and direct chill casting of large billets. In addition to its direct effect on mechanical properties such as strength, grain refinement is also used to reduce formation of defects such as hot tears, to control porosity, and to provide nucleation sites for recrystallisation dur- ing any subsequent thermo-mechanical processing [12]. The most potent grain refining addition known for magnesium alloys is zirconium [1]. The addition of a small amount of zirco- nium to the magnesium melt can readily be used to reduce the as-cast grain size by nearly two orders of magnitude as well as pro- ducing a nearly equiaxed and thus more uniform microstructure [2,3]. Zirconium is thus used as a grain refiner in most commercial magnesium alloys, providing they do not contain another addition that poisons the grain refining effect. The most important example of poisoning occurs in magnesium–aluminium alloys, where zirco- nium cannot be used for grain refinement because a stable com- pound is formed between zirconium and aluminium in the melt [2]. For this reason, to improve the strength of magnesium alloys grain refined by zirconium, an alternative major alloying is needed, and zinc is most commonly used [1]. In the most widely used commercial magnesium–zinc–zirconium (ZK) alloys, zinc is added at a high enough level to produce an age hardening effect by pre- cipitation on suitable heat treatment. The Mg–Zn system shows one of the highest age hardening responses of any magnesium alloy [4,5]. However, despite the fact that the volume fraction of precipitation that can occur in ZK60 (the most commonly used Mg–Zn–Zr variant) is comparable with that produced in age hard- enable aluminium alloys of the 2xxx and 7xxx series, the ageing response is an order of magnitude smaller [6,7]. As a result of the modest age hardening response of ZK alloys in comparison to equivalent aluminium alloys, the contribution of grain size strengthening is of increased importance [6]. Indeed, in contrast to the case of aluminium alloys, a T6 temper (solution treatment and age) is rarely applied to wrought ZK60 because the strength loss that occurs due to the increase in grain size during solution treatment overwhelms the strength increase asso- ciated with precipitation of the solute during ageing. http://dx.doi.org/10.1016/j.actamat.2015.05.012 1359-6462/Ó 2015 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). Corresponding author. E-mail address: [email protected] (J.D. Robson). Acta Materialia 95 (2015) 10–19 Contents lists available at ScienceDirect Acta Materialia journal homepage: www.elsevier.com/locate/actamat

Upload: others

Post on 11-Feb-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

Acta Materialia 95 (2015) 10–19

Contents lists available at ScienceDirect

Acta Materialia

journal homepage: www.elsevier .com/locate /actamat

The interaction of grain refinement and ageingin magnesium–zinc–zirconium (ZK) alloys

http://dx.doi.org/10.1016/j.actamat.2015.05.0121359-6462/� 2015 Acta Materialia Inc. Published by Elsevier Ltd.This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

⇑ Corresponding author.E-mail address: [email protected] (J.D. Robson).

J.D. Robson ⇑, C. Paa-RaiManchester Materials Science Centre, Grosvenor Street, Manchester M1 7HS, UK

a r t i c l e i n f o a b s t r a c t

Article history:Received 16 April 2015Revised 8 May 2015Accepted 9 May 2015Available online 31 May 2015

Keywords:PrecipitationStrengtheningMagnesium alloysModelling

Magnesium–zinc–zirconium (ZK) alloys are precipitation strengthened by the addition of zinc and grainrefined by zirconium. In ZK alloys these mechanisms interact since some Zn can be precipitated intoZn–Zr intermetallics. This reduces the fraction of age hardening precipitates. An experimental studyhas confirmed the presence of these two particle types in ZK60, a commercial Mg–Zn–Zr alloy. A modelhas been developed to consider precipitation and strengthening in ZK alloys accounting for this interac-tion. It has been shown that whilst Zr additions will increase strength by grain refinement this is balancedby a loss of age hardening response. It is predicted that if the same fine grain size could be achievedwithout Zr then a significant strength increase of around 15% would be expected. Therefore, in wroughtapplications where the final grain size is determined by recrystallisation, the use of Zr grain refiner in ZKalloys does not lead to optimum strengthening. The model was then applied to explore strategies toincrease strengthening in Mg–Zn alloys.� 2015 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article under the CC BY license

(http://creativecommons.org/licenses/by/4.0/).

1. Introduction

Grain refinement and age hardening (precipitation strengthen-ing) are two metallurgical phenomena that are exploited in magne-sium alloys to achieve the desired microstructure and mechanicalproperties. The refinement of grains in the as-cast condition gener-ally leads to an improvement in mechanical properties and struc-tural uniformity in most metals and alloys. The use of a grainrefining addition to promote the formation of a fine grain structureis particularly important for magnesium alloys produced underconditions where the solidification rate is not very rapid, such asencountered in sand casting, permanent mould casting, and directchill casting of large billets. In addition to its direct effect onmechanical properties such as strength, grain refinement is alsoused to reduce formation of defects such as hot tears, to controlporosity, and to provide nucleation sites for recrystallisation dur-ing any subsequent thermo-mechanical processing [12].

The most potent grain refining addition known for magnesiumalloys is zirconium [1]. The addition of a small amount of zirco-nium to the magnesium melt can readily be used to reduce theas-cast grain size by nearly two orders of magnitude as well as pro-ducing a nearly equiaxed and thus more uniform microstructure[2,3]. Zirconium is thus used as a grain refiner in most commercial

magnesium alloys, providing they do not contain another additionthat poisons the grain refining effect. The most important exampleof poisoning occurs in magnesium–aluminium alloys, where zirco-nium cannot be used for grain refinement because a stable com-pound is formed between zirconium and aluminium in the melt[2].

For this reason, to improve the strength of magnesium alloysgrain refined by zirconium, an alternative major alloying is needed,and zinc is most commonly used [1]. In the most widely usedcommercial magnesium–zinc–zirconium (ZK) alloys, zinc is addedat a high enough level to produce an age hardening effect by pre-cipitation on suitable heat treatment. The Mg–Zn system showsone of the highest age hardening responses of any magnesiumalloy [4,5]. However, despite the fact that the volume fraction ofprecipitation that can occur in ZK60 (the most commonly usedMg–Zn–Zr variant) is comparable with that produced in age hard-enable aluminium alloys of the 2xxx and 7xxx series, the ageingresponse is an order of magnitude smaller [6,7].

As a result of the modest age hardening response of ZK alloys incomparison to equivalent aluminium alloys, the contribution ofgrain size strengthening is of increased importance [6]. Indeed, incontrast to the case of aluminium alloys, a T6 temper (solutiontreatment and age) is rarely applied to wrought ZK60 becausethe strength loss that occurs due to the increase in grain sizeduring solution treatment overwhelms the strength increase asso-ciated with precipitation of the solute during ageing.

Page 2: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19 11

A further complication arises because the addition of zinc andzirconium cannot be considered in isolation due to an interactionbetween these elements. This is because there are a number ofcompounds that can form between them. Depending on the com-position, these compounds can form during solidification (in theliquid) or during subsequent heat treatment (in the solid). Theeffect of zinc concentration on the zirconium retained in solutionand the grain refining effect of this zirconium has been studiedpreviously by Hildebrand et al. and is discussed in detail later[3]. However, interaction between these elements will alsoinfluence the age hardening of the alloy, since zinc removed intozinc–zirconium compounds will no longer be available to formage hardening precipitates. This secondary effect has not beenexplicitly considered previously, but could have a significant effecton the age hardening capability of the alloy.

To improve the strength of Mg–Zn–Zr alloys therefore requiresa consideration of three objectives; 1. development and retentionof a fine grained structure through addition of grain refiner, recrys-tallisation (in the case of wrought products) and control of graingrowth, 2. maximising age hardening potential by ensuring theage hardening elements are not bound up in non-strengtheningphases, 3. refining precipitation by promoting particle nucleation.These aspects are not interdependent but are inter-related. Forexample, Zr addition will help to achieve objective 1 but will bedetrimental to objective 2. Maximising the available solute (objec-tive 2), which would usually be achieved by a solution treatment,will be detrimental to objective 1 if grain size during this processis not controlled.

In the present work, a model has been developed to simulateprecipitation in Mg–Zn–Zr alloys and predict the overall strength-ening of such alloys. This model has been used to help understandthe complex interactions identified above. The model has thenbeen applied to explore the potential to improve the age hardeningresponse of ZK alloys through optimisation of Zr level, control ofgrain size without the use of Zr, and refinement of precipitation.

0.35 0.4 0.45 0.530

40

50

60

70

80

90

100

Zr in solution / wt%

As c

ast g

rain

siz

e / μ

m

Experiment (Hildebrand et al.) Curve fit (0−0.5wt% Zr)

0 1 2 3 4Zinc addition / wt%

Fig. 1. Plot of limiting Zr concentration in solution and associated as-cast grain sizeas a function of added Zn level, data from Hildebrand et al. [3].

2. Experimental

Precipitation in the Mg–Zn system has been well documented[4,5,8,9], and it was not necessary to repeat this work. However,a small experimental study was performed to observe the interac-tion between the b01-MgZn strengthening phases and the Zn–Zrintermetallics in commercial ZK60 alloy (Mg-6 wt.% Zn-0.6 wt.%Zr). Cubes, with side length 20 mm were cut from the centre of acommercially extruded bar of ZK60 of diameter 110 mm producedwith an extrusion ratio of 10. These specimens were subject to atwo step solution treatment of 375 �C for 2 h followed by 2 h at500 �C under a protective argon atmosphere. This two step treat-ment is designed to maximise the amount of Zn in solution whilstavoiding incipient melting and similar treatments have been usedby previous researchers on ZK60 [10,11]. Ageing heat treatmentwas then performed at either 150 or 200 �C for times ranging from1 h to 384 h (16 days).

Specimens were prepared using standard metallographic tech-niques for hardness testing, which was performed with a Vickersmicrohardness testing machine on polished samples operated witha load of 300 g. Specimens for transmission electron microscopy(TEM) were prepared by mechanical punching of 3 mm diameterdiscs, which were then ground to a thickness of 50 lm. Afterdimpling, thin area was produced by ion milling with a Gatanion polisher operating at 4 kV with a tilt of 6�. TEM was performedusing a Philips CM200 and Technai T20 microscope operated at200 kV. Measurements of particle radius and length were madeby imaging along the h11 �20i or h10 �10i matrix zone axis. At least100 particles were measured for each ageing condition. The mean

aspect ratio was determined by dividing individual rod lengths bytheir diameter and finding the average for all ageing times.

3. The model

The goal of the model developed in this work is to provide acomplete prediction of the strengthening contributions in ZK alloysand hence predict the potential for strengthening in this systemwith optimised chemistry and processing. This involves also devel-oping a model for the formation of the b01-MgZn age hardeningprecipitates, accounting for the Zn removed into insoluble Zn–Zrparticles. The models for the individual contributions to strengthare described below.

3.1. Grain size strengthening

To calculate the contribution of grain boundaries to strengthen-ing it is first necessary to determine the grain size. In the as-castcondition, there will be a relationship between the grain size andlevel of Zr addition. In this case, an estimate of the grain size canbe made as a function of composition as described below. Inwrought products, the as-cast grain structure will usually bedestroyed by recrystallisation and grain growth processes. In thiscase, a simple prediction of grain size is not possible, and thisbecomes another input to the model derived from measurements.

Hildebrand et al. [3] have measured the grain size and thelimiting amount of zirconium in solution in a range of alloys withdifferent zinc levels. These data can be used to derive a simpleempirical curve fit to estimate the grain size in Mg–Zn–Zr alloysfor a given zirconium concentration. The data of Hildebrand et al.are replotted in this way in Fig. 1. A simple empirical (quadratic)fit to this data was used in the present work (as shown in Fig. 1).Note that the plot is scaled to highlight the composition rangesof most interest to this study; the curve fit was produced withadditional data at lower zirconium concentrations.

As Fig. 1 shows, the grain size decreases with increasing solubleZr concentration, and the soluble Zr concentration scales approxi-mately linearly with the level of Zn addition up to around 4 wt.%.Above this Zn level, the soluble Zr decreases due to intermetallicZn–Zr particle formation in the liquid metal [3] (not shown inthe plot but accounted for in the model). The interaction ofprocessing, composition, and grain refinement is highly complex,

Page 3: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

12 J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19

depending on the pre-heat temperature and stirring in the melt. Italso depends on the size of the Zr particles [12]. There is not yet acomplete model for grain refinement in magnesium that canaccount for all of these factors. Although more sophisticated mod-els do exist than the purely empirical approach used here [12],these models require calibration against experimental data andwould not therefore add extra predictive capability. Therefore,for the purposes of the present work, the simple empiricalapproach adopted here was sufficient to explore the interactionof Zn and Zr with regard to grain refinement.

Note that the grain refinement effect associated with Zr satu-rates at an addition of approximately 0.6 wt.% Zr, and further Zradditions do not provide any additional grain refinement [13]; thisis accounted for in the model by applying a limit to the minimumgrain size at Zr levels greater than 0.6 wt.%.

Once the grain size is known, the strength contribution fromgrain boundaries can be calculated using the Hall Petch method:

rgs ¼ kd�12 ð1Þ

where d is the grain size and k is the Hall Petch constant. Note, theintrinsic strength appears later, so is not included in this equation.For quenched Mg–Zn alloys, the value of k has been determined byChun et al. [14] and is reported in Table 1.

3.2. Interaction between zinc and zirconium

Only the Zn available in solution in the magnesium matrix cancontribute to precipitation during age hardening heat treatment.As will be shown later, even in the solution treated condition,the Zr present in ZK alloys will remove some of the Zn from thematrix into insoluble equilibrium Zn–Zr particles. A number ofsuch phases are possible, including ZnZr and Zn2Zr [3]. The moreZr present in the alloy, the greater the fraction of such phases,and thus the lower the effective Zn remaining in the matrix for age-ing. A further complication arises because in the as-cast condition,the Zr is not uniformly distributed, but is instead concentratedtowards the centre of the dendrites due to partitioning of Zr duringsolidification. Zr is a very slow diffuser in Mg, and so even afterhomogenisation it will remain non-uniform in its distribution,leading to the well documented Zr rich clouds seen towards thedendrite centres [15].

In the present work, a Scheil Gulliver solidification modelimplemented in the thermodynamic calculation software Pandatwith the MGDATA database (version 8) was used to calculate thedistribution of the alloying elements across a dendrite after

Table 1Constants and fitted parameters used by the precipitation and strengthening models.

Symbol Parameter Value Source

c Interfacial energy for nuclei 0.02 Jm-2 FittedN Nucleation site density

(relative)10�9 Fitted

D0 Pre-factor for Zn diffusion 1:05� 10�14 m2 s-1 [26]

QL Activation energy for Zndiffusion

125.8 kJ mol-1 [26]

l=t Length to diameter aspect ratio 10 Thiswork

r0 Intrinsic lattice strength 11 MPa [7]k Hall–Petch constant 0.35 MPa m1

2 [14]

C Solid solution constant 33 MPa [20]Eb Precipitate stiffness 36 GPa [27]G Matrix shear modulus 17.2 GPa [7]m Matrix poisson ratio 0.35 [7]b Burgers vector length 0.32 nm [7]M Taylor factor (as cast) 4.5 [19]M Taylor factor (extruded) 2.1 [19]

casting. Pandat was then used to calculate the amount of Zn thatwould be removed considering the local Zr level (at the homogeni-sation temperature) due to formation of Zn–Zr particles, which asshown later were of composition Zn2Zr. Zinc lost to both primaryZr particles (predicted to be present when sufficient Zr was added)and Zn2Zr particles that precipitate during homogenisation wereboth accounted for. The effective Zn thus calculated (i.e. the freeZn remaining in solution) was used as an input to the precipitationstrengthening model.

3.3. Precipitation strengthening

Precipitation strengthening in Mg–Zn alloys occurs through theformation of b01-MgZn, which form mainly as rod shaped particlesaligned parallel to the c-axis of the magnesium unit cell. To predictthe precipitation of these particles, a classical kinetic model wasused, based on the Kampmann and Wagner numerical (KWN)method [16]. Details of this model applied to continuous precipita-tion in AZ91 are given elsewhere [17]. In the present work, thismodel was modified to consider the different diffusing species con-trolling precipitation (Zn) and particle shape. It was assumed thatall precipitates form as c-axis rods [18] that grow with a fixedaspect ratio. The remaining unknown parameters in the modelare the effective interfacial energy and nucleation site density fac-tor. Although these parameters have a physical interpretation, herethey can be treated as empirical fitting factors; interfacial energydetermines the temperature sensitivity of the model whilst the sitedensity scales the nucleation at all temperatures. The fittingparameters were determined by tuning the model to experimentaldata, as discussed later. The full set of internal parameters used inthe model is shown in Table 1. Inputs to the model are the ageingtime and temperature and the effective zinc concentration.

The kinetic model provides a prediction of particle number den-sity and size for a given ageing treatment. These particles will pro-vide strengthening, and the contribution they make to strength iscalculated using an Orowan based hardening law. This assumesthat the particles are bowed rather than cut by dislocations; thisassumption has previously been demonstrated to be reasonablefor the precipitates in Mg–Zn alloys, at least in the peak agedand overaged conditions [5].

The Orowan model was adapted from a strengthening modeldeveloped for AZ91 by Hutchinson et al. [7]. The difference in ZKalloys is in the aspect ratio of the particles, but the methodologyand equations used to calculate the spacing between precipitatesare identical to those of basal plates (as formed in AZ91) [7], sincec-axis rods and basal plates are both cylinders of the same orienta-tion (but different aspect ratio). Following Hutchinson et al., it isassumed that basal slip controls yield, so that only strengtheningagainst basal dislocation motion is calculated [7]. This will bereasonable in the case of a randomly textured casting or weaklytextured wrought alloy, but may not be true in strongly texturedalloy (e.g. extrusion) loaded in an orientation in which the resolvedshear stress for basal slip is very low [9]. In the present model, theeffect of texture effects on the hardness evolution is accounted forby using an appropriate texture dependent value of the TaylorFactor to relate single crystal to polycrystalline behaviour, follow-ing Caceres et al. [19]. The Orowan equation gives the strengthcontribution due to bowing of precipitates as:

rOrowan ¼MGb

2pffiffiffiffiffiffiffiffiffiffiffiffi1� mp 1

kln

dA

r0

� �ð2Þ

where M is the Taylor factor, G is the shear modulus, b is theBurgers vector of the dislocation; m Poissons ratio; k the inter parti-cle spacing, and dA and r0 are respectively, the mean spacing andmean diameter of particles in the slip plane; r0 is the inner cut-off

Page 4: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19 13

radius of the dislocation taken equal to b. The calculation of theinter-particle spacing accounts for the preferred precipitate orienta-tion and particle shape, as described in detail elsewhere [7,18].

A key assumption of the Hutchinson et al. model (in commonwith most Orowan strengthening models) is that the particles arehomogeneously distributed in space, so that a knowledge of theaverage number density, size, and shape of the particles is suffi-cient to determine the inter particle spacing. However, it has beenknown that precipitation in the Mg–Zn system is heterogenous,and as will be demonstrated later this heterogeneity is exacerbatedby the addition of zirconium. It has previously been shown thattrue effective inter particle spacings in the Mg–Zn system can bearound 15% greater than that calculated assuming a homogeneousparticle distribution [5]. The homogeneous particle approximation,which is necessary to make the modelling problem tractable,therefore limits the accuracy of the model. However, the aim ofthe present work is not to provide a highly accurate prediction ofabsolute strengthening, but rather to explore the relative effecton strengthening of different alloy design strategies.

In addition to the Orowan strength due to precipitates there is acontribution from the backstress caused by the strain incompati-bility between unsheared precipitates embedded in a shearedmatrix. As shown later, this makes a minor contribution to thestrengthening in the present case since only very low strains areconsidered, but is accounted for following the same approximatemethod used by Hutchinson et al. [7].

rbs ¼ EbVbf � ð3Þ

where Eb is Young’s modulus of the precipitate, Vf is the volumefraction of precipitate, and � is the applied plastic strain.

The KWN precipitation model also gives the mean level of Znremaining in solution as an output, and this can be used tocalculate the solid solution strengthening contribution due to thiselement. Solid solution strengthening of Mg by Zn has been thesubject of numerous studies since it was identified in early workas an element that can provide strong solute strengthening againstbasal slip but can soften prismatic slip [20,21]. This is useful sinceit can potentially reduce the plastic anisotropy inherent in Mgcrystals and improve ductility [22]. Solid solution hardening ofMg by Zn is discussed in detail in work by Cáceres and Blake[22,23]. For the purpose of the present work, where it assumed thatbasal slip controls yield, it is sufficient to note that the solutehardening due to Zn can be well fitted to a power law relation ofthe form:

rss ¼ CX23Zn ð4Þ

(b(a)

200umED

Fig. 2. (a) Optical micrograph showing the grain structure of ZK60 in the solution tre

where XZn is the atomic fraction of Zn in solution, and C is a constant(Table 1).

3.4. Overal strengthening

To calculate the total strength after ageing, the contribution ofall strengthening mechanisms must be considered. FollowingHutchinson et al. [7], the total strength is expressed as a linearsummation of all the strength contributions, with a more sophisti-cated superposition model not justified given the approximationsused to calculate the individual contributions:

rtotal ¼ r0 þ rgs þ rss þ rbs þ rOrowan ð5Þ

where r0 is the intrinsic lattice resistance (taken as 11 MPa [7]) andthe other terms are as defined previously.

As discussed, the contribution of Hall–Petch strengthening willdepend on the grain size, which in turn will depend on the level ofZr addition. The amount of free Zn available for solid solution andprecipitation strengthening also depends on the Zr concentration.Solid solution strengthening and precipitation strengtheningclearly vary in the opposite sense as ageing proceeds; as solute islost from the matrix into precipitates during ageing, the solid solu-tion contribution decreases but the contribution from precipitationstrengthening increases.

4. Results and discussion

4.1. Microstructural observation

The grain structure of the ZK60 extrusion in the solution treatedcondition is shown in Fig. 2(a). Such a heterogeneous grain struc-ture is typical of large section commercial magnesium extrusions.The average grain size was measured as 35� 5 lm.

The measured evolution of hardness after ageing of ZK60 at 150and 200 �C is shown in Fig. 2(b). It can be seen that the hardeningresponse after ageing at 150 �C is greater than that at 200 �C, but inboth cases the increase in hardness is modest. For example, evenafter ageing to the peak (T6) condition at 150 �C, the hardness isincreased by only 40% over that in the solution treated condition.Contrast this with a high strength aluminium alloy (e.g. AA7050)where the hardness is increased by approximately 500% with aT6 heat treatment [24].

For reasons discussed later, the remainder of the study focussedon material aged at 200 �C. Example micrographs after ageing for10 h at this temperature (peak strength condition) are shown inFig. 3. The low magnification image (Fig. 3(a)) shows the presence

102

103

104

105

106

107

50

55

60

65

70

75

80150oC

200oC

Time / s

HV

)

ated condition. (b) Measured hardness response after ageing at 150 and 200 �C.

Page 5: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

1 um

Zn

Zr

(b)(a)

(c)

(d)i

ii

Fig. 3. Transmission electron microscopy of ZK60 aged at 200 �C for 10 h. (a) Low magnification image showing clusters of Zn2Zr particles (example arrowed). (b) Highmagnification image showing the two particle types present (i) Zn2Zr (ii) b01–MgZn. (c and d) EDX composition maps confirming the presence of Zn and Zr in the particlearrowed in (a).

50

T[C

]

0 2 4 6 8

w%(ZN)0 2 4 6 8

50

250

450

650

wt%(ZN)

T[C

]HCP+Zn2Zr

HCP+LIQUID+Zn2Zr

HCP+LIQUID+ZnZr

HCP+MGZN+Zn2Zr

HCP+MG7ZN3+Zn2Zr

HCP+ZnZr

HCP+LIQUID

HCP+Zn2Zr+ZnZr

HCP

ZK60

Fig. 4. Isopleth section of Mg–Zn–Zr phase diagram predicted using Pandat at afixed zirconium level of 0.6 wt.% Zr.

14 J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19

of two types of particle, as expected. The darker, globular particlescontain Zn and Zr; an example of one such particle is highlightedwith an arrow. An enlarged image of such a particle, and an EDXcomposition map confirming the presence of Zr and Zn, are shownin Fig. 3(b)–(d) respectively. Quantitative EDX analysis revealed aZn:Zr ratio close to 2:1 for these particles, this is consistent withthe expected Zn2Zr phase from thermodynamic calculations andprevious observations of such particles in Mg–Zn–Zr alloys [3].

It can be seen in Fig. 3(a) that these particles are very heteroge-neously distributed. This is to be expected, since they form in thehighly segregated as-cast microstructure, as discussed in detail inthe next section. Bands of finer needle shaped particles can alsobe seen in the microstructure, these were confirmed by diffractionanalysis as b01-MgZn, and no Zr was detected in them. Note thatwhere there is a dense cloud of Zn2Zr (in the upper left ofFig. 3(a)), no b01-MgZn is observed. This is presumably due to thelocal depletion of matrix Zn in these regions into Zn2Zr prior toageing, although accurate quantification of the matrix chemistrywas not possible with EDX due to the influence of the particles.

Fig. 3(b) shows a higher magnification image where the twotypes of particle are highlighted (i) Zn2Zr and (ii) b01-MgZn. Again,it can be seen that there is a region around the Zn2Zr particle whereno b01-MgZn is observed. The Zn2Zr particles themselves are toolarge and widely spaced to make a significant direct contributionto strengthening. Although it was not accurately quantified, theTEM observations show these particles are typically in the range100–200 nm in diameter with inter-particle spacings between100 nm and 1 lm (the large variation being due to the heterogene-ity in the spatial distribution of these particles). These particles willnot make a large direct contribution to the strength of the alloysince their spacing is an order of magnitude greater than that ofthe age hardening b01-MgZn rods.

The TEM observations support the hypothesis that the interac-tion of Zr and Zn in commercial ZK60 leads to a reduced precipita-tion of strengthening b01-MgZn. They show that in practice, this is ahighly heterogeneous phenomena, with local variations in Zr lead-ing to local differences in the fraction of Zn2Zr and hence depletionof Zn from the matrix.

4.2. Thermodynamic calculations

A calculated isopleth section through the ternary Mg–Zn–Zrdiagram at 0.6 wt.% Zr (a typical level used in commercial ZK60)is shown in Fig. 4(a). The isopleth section shows the effect of add-ing Zr on the solubility of Zn and on the ability to fully solutionisethe alloy. It can be seen that for any Zn concentration is greater

than 0.5 wt.%, there is no temperature where it is possible to fullysolutionise all of the Zn into the HCP Mg matrix when 0.6 wt.% Zr isalso present. For the Zn concentration used in ZK60 (6 wt.%), it canbe seen that Zn2Zr is predicted to be the stable phase at thehomogenisation temperature.

Although such equilibrium calculations are useful, in practicethe distribution of Zr is highly non-uniform due to segregation dur-ing casting. The predicted distribution of Zr across a dendrite in theMg matrix after casting is shown in Fig. 4(b) for the case of ZK60(0.6 wt.% Zr). This can be used to calculate the local Zn that willbe removed by precipitation of Zn2Zr, which will occur directlyfrom the as-cast structure. Note that precipitation will of Zn2Zr willonly occur when the local zirconium concentration is sufficient toexceed the solubility limit, as indicated in Fig. 4(b). The locallydepleted Zn is assumed to redistribute uniformly across the den-drite due to diffusion at the solution treatment temperature, andthus the local depletion can be used to calculate an average globaldepletion (marked on Fig. 4(b)). Finally, the Zn lost by transforma-tion of primary Zr into Zn2Zr must also be accounted for; thismakes a relatively small contribution to the overall Zn depletionfor the (typical) Zr level used in this calculation (0.6 wt.%) as shownin Fig. 4(b).

It can be seen that for ZK60 and the conditions used in the pre-sent study, there is a depletion of Zn of 0.39 at.% due to precipita-tion into Zn2Zr; this corresponds to approximately 17% of the Znadded to the alloy being locked up in coarse non-strengtheningZn2Zr particles. As shown later, this has a significant effect on thepotential age hardening response of the alloy.

Page 6: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

100 1020

0.005

0.01

0.015

0.02

0.025

Time / h

Volu

me

Frac

tion

100 1020

50

100

150

200

250

300

350

Time / h

Rad

ius

/ nm

100 1020

200

400

600

800

1000

Time / h

Spac

ing

/ nm

100 10245

50

55

60

65

70

75

80

Time / h H

ardn

ess

/ VH

N

Exp (Jain)Model

(c) (d)

(a) (b)

Fig. 5. Model predictions of (a) precipitate volume fraction (b) average precipitate radius (c) average precipitate spacing and (d) overall hardness evolution compared withexperimental data from Jain et al. [9], Mg-6 wt.% (2.3 at.%) Zn, aged at 200 �C.

J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19 15

4.3. Model validation

To validate the model it was first applied to the simpler case ofpredicting precipitation and strengthening in binary Mg–Zn alloysbefore considering the extra complexity associated with the addi-tion of Zr. The model requires a large number of input parameters,as shown in Table 1, each of which can influence the accuracy ofthe predictions. Most of these parameters can be extracted fromthe literature, but there are two critical model inputs that mustbe found by fitting. These inputs are the effective interfacial energyof the precipitates and effective nucleation site density. As dis-cussed elsewhere, a unique set of these parameters can be foundby matching the KWN kinetic model predictions to experimentaldata from several temperatures [17]. Two datasets from the litera-ture that contain detailed information about the size, spacing, andvolume fraction of precipitates and their effects on hardness evolu-tion were used to test the model [9,8]. These two datasets corre-spond to alloys with two different compositions aged at twodifferent temperatures (150 and 200 �C). The effective interfacialenergy and effective nucleation site density parameters weretuned so that a single parameter set gave the best overall fit tothese data. This optimised parameter set was then used for all fur-ther calculations.

Fig. 5 shows the best fit obtained between model predictionsand the reported measurements for a Mg-6 wt.% (2.3 at.%) Zn stud-ied by Jain et al. [9] after ageing at 200�C. It can be seen that themodel predictions are in reasonable agreement with the measure-ments. In particular, the inter particle spacing, which is the mostcritical parameter in calculating the strengthening due to the par-ticles, is accurately predicted. The volume fractions measured byJain et al. are lower than what would be expected; for examplethe terminal volume fraction they measured (0.018) is consider-ably less than the equilibrium volume fraction (0.22), which wouldbe expected to be reached at long ageing times. Jain et al. used TEM

to determine their parameters, and the measured volume fractionwith this method is susceptible to error [8]. It is suggested that thevolume fraction measured by Jain et al. is a slight underestimate ofthe true fraction, and when this is considered the model predictedfraction is in very good agreement with the measurement. Thehardness evolution is also well captured, especially consideringthe approximations necessary in the model. The peak hardness isslightly over-estimated, at least one major contributing factor tothis is the assumption in the model that the precipitates are uni-formly distributed in space, which is not the case in practice as pre-viously discussed.

Fig. 6 shows the fit obtained with the same parameters for a Mg-3.4 at.% Zn alloy studied by Rosalie and Pauw [8] after ageing at150 �C. The data from Rosalie and Pauw were determined usingSmall Angle X-ray Scattering (SAXS) rather than TEM. Both tech-niques have their advantages and disadvantages, and will givesomewhat different results. In this case, the model fit is less good.However, given the approximations inherent in the model anduncertainties in the input parameters, the agreement is consideredto be reasonable. In particular, the volume fraction is always within0.01, the particle radius is within 2 nm, and the hardness is within20 HVN of that measured. The overestimate of hardness at this tem-perature is likely to be due to the assumption in the model of a spa-tially uniform precipitate distribution as at 200 �C, but it also seemslikely that at this temperature, where particles are smaller, Orowanlooping may not be the only dislocation bypass mechanism. It ismore likely that at this ageing temperature, a significant fractionof the rods will be below the critical size for cutting by dislocations.This mechanism is not considered in the model, so that when a sig-nificant fraction of the particles are cut, their contribution tostrength and hardness will be over estimated. Since this remainsa topic of uncertainty, all subsequent calculations are performedfor ageing temperatures of 200 �C, where experiments have provedthat bowing is the dominant bypass mechanism [9].

Page 7: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

100 1020

0.005

0.01

0.015

0.02

0.025

0.03

0.035

TIme / h

Volu

me

Frac

tion

100 1020

1

2

3

4

5

6

7

8

Time / h R

adiu

s / n

m

100 10255

60

65

70

75

80

85

90

Time / h

Har

dnes

s / H

VN

ModelExp (Rosalie)

(a) (b) (c)

Fig. 6. Model predictions of (a) precipitate volume fraction (b) average precipitate radius and (c) overall hardness evolution compared with experimental data from Rosalieand Pauw [8], Mg-3.4 at.% Zn, aged at 150 �C.

16 J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19

4.4. Application to Mg–Zn–Zr

Having validated the model for the simpler case of binary Mg–Zn alloys, it was then applied to the more complex commercialMg–Zn–Zr case. Fig. 7 shows the predictions of the model (particlelength and hardness) compared with the measurements made inthe present work for ZK60. As explained in the section describingthe model, in these calculations the effect of zirconium in remov-ing some of the zinc into insoluble Zn–Zr precipitates is accountedfor. It can be seen that the model captures the measurements rea-sonably and within the error limits of the experimental data with-out any further recalibration. The large errors bars (standarddeviation) on the experimental data points reflect the large scatterin particle sizes measured due to the heterogeneous nature of themicrostructure as previously demonstrated.

The model can be interrogated further to plot separately thepredicted contributions to the overall strength (for the case where

100 1020

5

10

15

20

25

Time / h

Equi

vale

nt ra

dius

/ nm

ExpModel(a)

Fig. 7. Model predictions of (a) precipitate radius and (b) overall hardness evolution

the as-cast grain size is preserved). Fig. 8 shows the predicted con-tributions to strengthening for Z6 and ZK60 (i.e. without and with0.6 wt.% zirconium respectively) during ageing at 200 �C. The grainsize and Orowan contributions to the strength are clearly the dom-inant factors. Without zirconium, the grain size is larger due to thelack of grain refinement so the grain size contribution to strength isless. However, it is predicted that this is more than compensatedfor by the increased Orowan hardening due to the lack of Zn lossinto Zn2Zr. It is noteworthy that in the binary Z6 alloy, it is pre-dicted that at peak strength the Orowan contribution from parti-cles is more than twice that of the grain size strengthening,whereas in ZK60 it is only 1.25 times greater. This emphasisesthe importance of Hall Petch strengthening to the total strengthof age hardenable magnesium alloys [6]. This is in contrast to highstrength aluminium alloys (e.g. AA7050) where the combinedeffect of Hall Petch and solid solution strengthening is only 20%of the strength that can be achieved in the peak aged condition

100 10245

50

55

60

65

70

Time / h

Har

dnes

s / V

HN

(b)

compared with experimental data from the present study, ZK60, aged at 200 �C.

Page 8: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

0.01 1 1000

50

100

150

200

250

300

Time / h

Stre

ngth

con

tribu

tion

/ MPa

0.01 1 1000

50

100

150

200

250

300

Time / h

Stre

ngth

con

tribu

tion

/ MPa

TotalGrainSolutionOrowanBackstress

Mg−6wt%Zn

Mg−6Zn−0.6 wt%Zr

(a) (b)

Fig. 8. Predicted contributions to overall strength on ageing at 200�C for (a) Mg-6 wt.% Zn, (b) Mg–6Zn–0.6 wt.% Zr (ZK60). Note that the grain size contribution includes theintrinsic strength (11 MPa).

J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19 17

[24]. Obtaining and retaining a fine grain size during processing istherefore particularly critical to obtaining good strength levels inmagnesium.

4.5. Effect of zirconium

The model was then used to explore the effect of varying zirco-nium concentration on the precipitation and strengthening in ZKmagnesium alloys. Zirconium was varied between 0 and 0.8 wt.%.Commercially, Zr is typically added at the level of around0.6 wt.%, since it has been shown that further additions do not leadto further grain refinement [1].

Fig. 9(a) shows the calculated volume fraction of strengtheningprecipitates (b01-MgZn) and inter particle spacing in the peak aged(maximum strength) condition after ageing at 200 �C. As the Zrconcentration increases, the volume fraction of age hardening pre-cipitates decreases and the spacing between precipitates increasesdue to Zn lost to Zn2Zr, as already discussed.

0

0.005

0.01

0.015

0.02

0.025

Zirconium addition / wt %

Volu

me

fract

ion

0 0.2 0.4 0.6 0.80

0.05

0.1

0.15

0.2

0.25

Volume fraction Spacing

(a)

Fig. 9. Predicted effect of varying Zr level in a Mg-6 wt.% Zn alloy on (a) precipitate volustrength condition at 200 �C. Note that the grain size contribution includes the intrinsic

Fig. 9(b) shows the contribution of individual strengtheningmechanisms to the overall strength. As expected, the Orowan con-tribution due to strengthening precipitates decreases as Zr concen-tration is increased. However, most of this loss is compensated forby the grain refinement associated with the Zr addition, up to a Zrlevel of 0.6 wt.%. Above this limit, any additional zirconium addeddoes not have a beneficial effect in increasing grain boundarystrengthening but does still have a detrimental effect in reducingthe age hardening potential due to zinc removal. Therefore, a rapiddrop in overall strength is predicted.

This analysis applies when material is used withoutthermo-mechanical processing (TMP), so that the grain sizeobtained after casting is retained in the final product. In wroughtZK60, the grain size will also be influenced by the imposed strain,temperature and the influence of dynamic and static recrystallisa-tion. In this case, the grain size is not be directly related to theas-cast grain size, and there will not be a direct compensatingeffect due to the competing effects of zirconium on grain size

Spac

ing

/ μm

0 0.2 0.4 0.6 0.80

50

100

150

200

250

300

350

Zirconium addition / wt%

Stre

ngth

/ M

Pa

GS+ intrinsicSoluteOrowanBackstress

(b)

me fraction and spacing (b) contributions to strengthening after ageing to the peakstrength (11 MPa).

Page 9: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

0

0.01

0.02

Time / h

Volu

me

fract

ion

10−1 100 101 102

Spac

ing

/ nm

0

500

1000

10−1 100 101 10245

50

55

60

65

70

75

Time / h

Har

dnes

s / V

HN

GR no Zr GR with Zr

(a) (b)

Vf no Zr

Vf with Zr

Spacingwith Zr

Spacingno Zr

Fig. 10. Predicted evolution of (a) precipitate volume fraction and spacing, (b) total hardness for a Mg-6 wt.% Zn alloy with a fixed grain size of 10 lm aged at 200 �C with andwithout 0.6 wt.% Zr addition.

0

0.005

0.01

0.015

0.02

0.025

0.03

Volu

me

fract

ion

10−1 100 101 1020

Time / h

0

0.05

0.1

0.15

0.2

0.25

Spac

ing

/ μm

10−1 100 101 10245

50

55

60

65

70

75

80

85

90

Time / h

Har

dnes

s / V

HN

Standard γReduced γ

spacing

Vf

(a) (b)

gs = 10 μmgs = 10 μm

Fig. 11. Predicted effect of a 10% reduction in the interfacial energy of b01-MgZn precipitates on (a) precipitate volume fraction and spacing (b) hardness evolution for a grainsize of 10 lm aged at 200 �C.

18 J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19

and age hardening potential. This effect is discussed in detail in thenext section.

4.6. Potential strength in wrought products

It has been demonstrated that Zr additions have a detrimentaleffect on the age hardening response of Mg–Zn alloys. If a fine grainsize could be achieved without the use of Zr, then there would bean advantage in age hardening potential. In wrought products,grain refinement can be achieved by TMP to induce recrystallisa-tion under controlled conditions of temperature, strain rate, andstrain. The fine grain structure obtained can be stabilised by suit-able pinning particles (dispersoids); the use of Mn additions in thisrole in Mg has been explored previously [25].

The model was applied to consider the case where a grain sizeof 10lm was achieved by TMP, with and without Zr. This grain sizeis readily attained by conventional hot rolling or extrusion of mag-nesium alloys, exploiting the tendency of the material to undergodynamic recrystallisation (DRX). Fig. 10 shows the precipitate vol-ume fraction and spacing predicted with and without Zr additions

(0 wt.% Zr and 0.6 wt.% Zr) for a fixed grain size of 10 lm. Asdiscussed previously, eliminating Zr results in a larger b01-MgZnvolume fraction and smaller interparticle spacing. This translatesto a increase in predicted maximum hardness from 62 to 71 VHN(a 15% increase); the same percentage increase in strength is pre-dicted. There is thus a significant incentive to eliminate Zr toimprove hardness and strength, providing the same fine grainedstructure can be obtained by an alternative route (e.g. TMP andrecrystallisation control).

4.7. Effect of promoting nucleation

One reason for the relatively poor age hardening response ofmagnesium alloys compared to aluminium alloys is the relativelylarge precipitate size and spacing that is typically attained afterageing of magnesium alloys. To address this issue, a number ofattempts have been made to promote precipitate nucleation inmagnesium, and thus produce a finer and more closely spaced par-ticle dispersion. Attempts to achieve this in Mg–Zn magnesiumalloys have included pre deforming the material [4,5], or adding

Page 10: The interaction of grain refinement and ageing in ... · The interaction of grain refinement and ageing in magnesium–zinc–zirconium (ZK) alloys J.D. Robson⇑, C. Paa-Rai Manchester

J.D. Robson, C. Paa-Rai / Acta Materialia 95 (2015) 10–19 19

a dopant to decrease the energy barrier to nucleation, such as Ca[6].

It has been shown that even though precipitation in Mg–Zndoes occur heterogeneously on dislocations, pre-deformation doesnot lead to a significant enhancement of ageing response when slipcontrols yield [5]. The reason is that even at relatively low ageingtemperatures (e.g. 150 �C) recovery processes are rapid and thedeformation introduced defects are annealed out at a shorter timethan the incubation time for nucleation [5].

Micro-alloying to reduce the energy barrier to nucleation(for example, by reducing the interfacial energy of the nuclei)and thus enhance the nucleation rate appears a more promisingapproach to improve age hardening response. Microalloying forthis purpose is widely used in aluminium alloys. The potential tofind suitable micro-alloying additions to achieve this goal inmagnesium is discussed in detail elsewhere [6].

The model developed here was applied to predict what effect areduction in interfacial energy would have on precipitation andstrengthening in the Mg–Zn system. The predictions for a reduc-tion in interfacial energy of 10% are shown in Fig. 11. The reducedinterfacial energy clearly leads to faster kinetics due mainly to anincrease in nucleation rate. This, in turn, leads to a smallerpredicted inter-particle spacing (Fig. 11(a)). The effect of this onthe predicted hardness evolution is to produce a 15% increase inpredicted peak hardness along with a reduced ageing timerequired to reach this hardness. These predictions show the poten-tial benefits that may be achieved by even a modest reduction ininterfacial energy produced (for example) by micro-alloying.

5. Conclusions

In the present work, a new model has been developed to predictprecipitation and strengthening in Mg–Zn–Zr (ZK) alloys. Themodel has been applied to understand the age hardening responseof ZK alloys and explore the interaction between Zn and Zr. Thefollowing conclusions can be drawn from this work

1. In common with other age hardenable magnesium alloys, theage hardening response of Mg–Zn–Zr alloys is poor comparedwith equivalent aluminium alloys. It is predicted that in ZK60,the most widely used commercial Mg–Zn–Zr alloy, age harden-ing only provides an addition to strength similar to thatprovided by the sum of grain size and solution strengthening.In high strength aluminium alloys, age hardening typically pro-vides over 5 times the strengthening of other mechanisms.

2. In addition to the effect of precipitate orientation and limitednucleation, which have been considered by previous research-ers, it is shown in the present work that the presence of Zr inMg–Zn–Zr alloys reduces the age hardening response due tothe formation of stable Zn2Zr intermetallics. These particlesremove some of the free Zn from solution. This effect is pre-dicted to lead to a strength reduction in ZK60 of around 15%at an ageing temperature of 200 �C.

3. The benefits from increased grain size strengthening due tograin refinement by Zr additions are predicted to be balancedby the reduced age hardening response due to the loss of Znfrom solution. In wrought products, where the final grain sizeis controlled by recrystallisation during thermomechanial pro-cessing (TMP) and is different from the as-cast grain size, thereis thus potential benefit to be obtained by eliminating Zr andcontrolling grain size through TMP.

4. The poor age hardening response of ZK60 alloys could beimproved by addition of micro alloying elements that promotenucleation of additional particles. It is predicted that even if

such an element has only a modest (e.g. 10%) reduction onthe interfacial energy of the precipitates, a significant (15%)increase in strength would be expected.

Acknowledgment

The authors are grateful to Magnesium Elektron UK for provi-sion of the ZK60 alloy used in this study. Thanks to EPSRCLATEST-2 programme Grant (EP/H020047/1) for financial supportfor this work.

References

[1] B.L. Mordike, Creep-resistant magnesium alloys, Mater. Sci. Eng. A 324 (2002)103–112.

[2] D. StJohn, M.A. Qian, M.A. Easton, P. Cao, Z. Hildebrand, Grain refinement ofmagnesium alloys, Metall. Mater. Trans. A 36 (2005) 1669–1679.

[3] Z. Hildebrand, M. Qian, D. StJohn, M. Frost, Influence of zinc on the solublezirconium content in magnesium and the subsequent grain refinement byzirconium, Magnes. Technol. (2004) 241–245.

[4] J.B. Clark, Transmission electron microscopy study of age hardenining in a Mg-5wt% Zn alloy, Acta Metall. 13 (1965) 1281–1289.

[5] J.M. Rosalie, H. Somekawa, A. Singh, T. Mukai, The effect of size anddistribution of rod-shaped precipitates on the strength and ductility of aMg–Zn alloy, Mater. Sci. Eng. A 539 (2012) 230–237.

[6] C.L. Mendis, K. Oh-ishi, Y. Kawamura, T. Honma, S. Kamado, K. Hono,Precipitation-hardenable Mg-2.4Zn-0.1Ag-0.1Ca-0.16Zr (at.%) wroughtmagnesium alloy, Acta Mater. 57 (2009) 749–760.

[7] C.R. Hutchinson, J.F. Nie, S. Gorsse, Modeling the precipitation processes andstrengthening mechanisms in a Mg–Al–(Zn) AZ91 alloy, Metall. Mater. Trans. A36 (2005) 2093–2105.

[8] J.M. Rosalie, B.R. Pauw, Form-free size distributions from complementarystereological tem/saxs on precipitates in a Mg–Zn alloy, Acta Mater. 66 (2014)150–162.

[9] J. Jain, P. Cizek, W.J. Poole, M.R. Barnett, Precipitate characteristics and theireffect on the prismatic-slip-dominated deformation behavior of an Mg–6Znalloy, Acta Mater. 61 (2013) 4091–4102.

[10] J.H. Cho, Y.M. Jin, H.W. Kim, S.B. Kang, Microstructure and mechanicalproperties of ZK60 alloy sheets during aging, Mater. Sci. Forum 558–559(2007) 159–164.

[11] Y. Li, Z. Zhang, Y. Xue, Influence of aging on microstructure and mechanicalproperties of AZ80 and ZK60 magnesium alloys, Trans. Nonferrous Met. Soc.China 21 (2011) 739–744.

[12] M. Sun, M.A. Easton, D.H. StJohn, G. Wu, T.B. Abbot, W. Ding, Grain refinementof magnesium alloys by Mg–Zr master alloys: the role of alloy chemistry andzr particle number density, Adv. Eng. Mater. 15 (2012) 373–378.

[13] H.E. Friedrich, B.L. Mordike, Magnesium Technology, Springer Verlag, Berlin,2006.

[14] J.S. Chun, J.G. Byrne, A. Bornemann, The inhibition of deformation twinning byprecipitates in a magnesium-zinc alloy, Philos. Mag. 20 (1969) 291–300.

[15] M. Qian, D.H. StJohn, M.T. Frost, Characteristic zirconium-rich coringstructures in Mg–Zr alloys, Scr. Mater. 46 (2002) 649–654.

[16] R. Kampmann, R. Wagner, P.W. Voorhees, Homogeneous second-phaseprecipitation, Mater. Sci. Technol. 5 (1991) 215–298.

[17] J.D. Robson, Modeling competitive continuous and discontinuousprecipitation, Acta Mater. 61 (2013) 7781–7790.

[18] J.F. Nie, Precipitation hardening in magnesium alloys, Metall. Mater. Trans. A43 (2012) 3891–3939.

[19] C.H. Caceres, P. Lukac, Strain hardening behaviour and the Taylor Factor ofpure magnesium, Philos. Mag. 88 (2008) 977–989.

[20] A. Akhtar, E. Teghtsoonian, Solid solution strengthening of magnesium singlecrystals-i alloy behaviour in basal slip, Acta Metall. 17 (1969) 1351–1356.

[21] A. Akhtar, E. Teghtsoonian, Solid solution strengthening of magnesium singlecrystals-ii the effect of solute on the ease of prismatic slip, Acta Metall. 17(1969) 1351–1356.

[22] A.H. Blake, C.H. Cáceres, Solid-solution hardening and softening of Mg–Znalloys, Mater. Sci. Eng. A 483–484 (2008) 161–163.

[23] C.H. Cáceres, A. Blake, The strength of concentrated Mg–Zn solid solutions,Phys. Status Solidi 194 (2002) 147–158.

[24] R.C. Dorward, Precipitate coarsening during overaging of Al–Zn–Mg–Cu alloy,Mater. Sci. Tech. 15 (1999) 1133–1138.

[25] J.D. Robson, D.T. Henry, B. Davis, Particle effects on recrystallization inmagnesium-manganese alloys: Particle pinning, Mater. Sci. Eng. A 528 (2011)4239–4247.

[26] J. Cermák, I. Stloukal, Diffusion of zinc in two-phase Mg–Al alloy, Defect Diffus.Forum 263 (2007) 165–170.

[27] P. Mao, B. Yu, Z. Liu, F. Wang, Y. Ju, First-principles calculations of structural,elastic and electronic properties of AB2 type intermetallics in Mg-Zn–Ca–Cualloy, J. Magn. Alloys 1 (2013) 256–262.