• Keine Ergebnisse gefunden

95

8. Summary

Solute clustering and precipitation during multi-stage ageing in a commercial AA6014 alloy are systematically investigated in this study, covering 1) the role of pre-ageing (PA) in multi-stage ageing treatments, 2) the influence of quench rate on multi-multi-stage ageing treatments, 3) the clustering and precipitation behaviour during linear heating, all by means of various characterisation methods including hardness testing, tranmission electron microscopy, electrical resistivity measurements, thermoanalysis, and positron lifetime measurements.

Conclusions corresponding to each specific study are stated at the end of chapters 4 – 7. Here, the most general findings pointed out in various chapters are summarised:

1) Direct vacancy annihilation alone cannot explain the excess vacancy evolution during natural ageing (NA or NSA). This has been demonstrated by the NA kinetics after various quenches or by the NSA kinetics after various PA treatments. Other mechanisms such as vacancy clustering and vacancy trapping by clusters must play a role in the reduction of vacancies.

2) During multi-stage ageing, clusters formed in early stages play a very important role in later stages, not only in a direct manner by transforming to precipitates, causing e.g.

different PB hardening effects after PA at different temperatures but also indirectly by influencing the vacancy evolution by vacancy trapping, which regulates the mobile vacancy site fraction by repartitioning the total vacancies (quick process) and buffering vacancy annihilation and generation (slow process).

Besides, the current study shows a way to improve industrial PA strategies and demonstrates that the behaviour of samples during multi-stage ageing under simulated industrial conditions is similar to that of samples processed by water quenches with just slightly inferior PB hardening. Industry might benefit from the current study to design better processes and improve the alloy properties.

96

Appendix A

Supplemental material to the effect of PA to NSA and PB hardening

A1. Additional hardness data

Fig. SA1 – SA3 show the hardening of the alloy during NSA and after subsequent PB for alloys that have been pre-aged under different conditions.

The curves in Fig. SA1a–h are analogous to those in Fig. 4.3a,e, just that PA conditions differ.

Fig. SA1i-n correspond to Fig. 3c,g.

Appendix A

97

98

Fig. SA1. (a, c, e, g) Hardening curves during NSA after PA at various temperatures for various times.

(b, d, f, h) Hardnesses after additional PB. (i, k, m) Comparison of NSA after PA to other hardnesses than those given in Fig. 4.3(c) of Chap. 4 (60 HBW), (j, l, n) displays hardnesses after additional PB.

Fig. SA2 displays hardness after PA and NSA as a function of PA hardness, A continuous evolution from short NSA time to long NSA time is seen. NSA hardness increases first in the low PA hardness regime and at high PA temperature regime.

Fig. SA3 corresponds to Fig.SA2, just that a final PB was carried out. The final PB hardness changes most after the first week of NSA in the low PA hardness regime.

Appendix A

99 Fig. SA2. Hardness after NSA as a function of PA hardness and temperature. Different graphs correspond to different NSA times. f) is identical to Fig. 4.3d in Chap. 4.

100

Fig. SA3. Similar to Fig. SA2 but with additional PB. f) is identical to Fig. 4.3h in Chap. 4.

A2. Additional DSC measurements

Fig. SA4a – c shows DSC traces of alloys aged at 100 °C for 60 min, 240 min, and pre-aged at 120 °C for 60 min and then NSA for various times. The trends of the DSC curve evolution for increasing NSA time are similar, just that the influence of the same NSA time on the DSC curve is dependent on PA condition. The heat content for clustering during ageing (PA and/or NSA) can be estimated by ∆𝐻ageing= 𝐻AQ− 𝐻after ageing, where 𝐻 is the

Appendix A

101 integrated DSC heat flow from 𝑇𝑎 to 𝑇𝑒, 𝐻 = ∫ 𝑑𝑄

𝑑𝑡𝑑𝑇

𝑇𝑒

𝑇𝑎 . 𝑇𝑎 is the temperature before the first reaction begins, and 𝑇𝑒 the temperature at which equilibrium is reached, e.g. above the solution heat treatment temperature. However, as pointed out by Esmaeili et al. [19, 144], due to a high baseline drift in the higher temperature regime, a lower 𝑇𝑒 can also be chosen as long as all relevant precipitation/dissolution peaks are covered within the temperature range. We choose 𝑇𝑎 = 50 °C and 𝑇𝑒 = 315 °C because DSC curves before 𝑇𝑎 and after 𝑇𝑒 are converged. 𝐻AQ is constant, thus 𝐻after ageing = ∆𝐻ageing+ 𝐻AQ. Fig. 4.8 shows hardness plotted as a function of the DSC integral for samples pre-aged and subsequently naturally aged (in cases where DSC and hardness were both measured). A linear relationship is seen although with some fluctuations possibly due to different samples and possibly some baseline drift. If the heat signal is assumed to be proportional to the precipitated volume the latter is also proportional to hardness after NSA.

(a)

102

(b)

(c)

Fig. SA4. DSC traces of samples PA a) at 100 °C for 60 min, (b) 240 min, (c) PA at 120 °C for 60 min with and without ensuing NSA.

Appendix A

103 A3. Retardation factor from resistivity measurements

Retardation factors were determined using hardness curves and collapsing them into one master curve by normalising the NSA time with Θ−1. As electrical resistivity is more sensitive to clustering than hardness, it might detect earlier changes during NSA. Fig. SA5 shows the data of Fig. 4.4d on a linear time scale. A linear increase is found at the beginning of NSA with almost no incubation time. Thus the retardation factor can be calculated from the slope of the initial resistivity increase [71]. Here relative retardation of PA 80/480 to PA 160/10 is 12, which is larger than the value obtained from hardness (ℛθ ≈ 3) but still smaller than the ratio between PA equilibrium vacancy site fractions (ℛc ≈ 58). This underlines the importance of PA clusters in determining the mobile vacancy site fraction and strengthen the arguments in Sec. 4.2.2.2. The main conclusions remain the same.

(a) (b)

Fig. SA5. (a) Same as Fig. 4.4d but plotted with linear NSA time. (b) the initial linear increase stage of PA 80/480 and PA 160/10. Curves in (b) have been smoothed by filtering artefact data points (sudden drops [145]) and averaging every 5 points.

A4. Possible other cluster strengthening mechanism

PA clusters formed at different temperatures but yielding the same hardness were assumed in Sec. 4.2.2 to consume the same fraction of solute supersaturation. Now we discuss if clusters formed at different temperatures have a different strengthening effect. The ‘Weak obstacle’

model [34, 146, 147] assumes a relationship between hardness increase and fraction of solute clusters, ∆ℋ ∝ √𝛼 ∙ √𝑟 with 𝑟 the radius of solute clusters. Since clusters formed at higher temperature have on average a larger radius [53], a smaller fraction of solutes is consumed when the same hardness is reached by PA at higher temperature. As a consequence, the

104

retardation factor ratio will be smaller than the one following from Fig. 4.10 if the same solute fraction has to be consumed at different PA temperatures. This does not affect the main conclusions of the work.

A5. Fitting hardness data by various kinetic models

We attempted to obtain kinetic parameters by fitting the hardness data during NSA, ℋ(𝑡), by two known functions. We present them here although none of these attempts yielded consistent kinetic parameters to illustrate the difficulties involved. The problems encountered are:

 The function does not represent the data well

 Representation is good but the use of too many parameters makes usage of function questionable.

Function parameters function ℋ(𝑡) =

1 Single Avrami [148] 0, ℋ1, 𝑘, 𝑛 0+ (ℋ1− ℋ0)[1 − 𝑒−(𝑘𝑡)𝑛] 2 Starink-Zahra [117] 0, ℋ1, 𝑘, 𝑛, 𝜂 0+ (ℋ1− ℋ0) {1 − [(𝑘𝑡)𝑛

𝜂 + 1]−𝜂}

Function #1 does not yield satisfactory fit results and the values for the Avrami-coefficient 𝑛 are unrealistically low in some cases (𝑛 = 0.5 for 10 min NA), see Fig. SA6a. For longer PA times 𝑛 ≈ 1 is found. The conditions for JMAK are not fulfilled in our case because the vacancy fraction is continuously decreasing, which is why use of the JMAK model is questionable.

By enforcing 𝑛 = 1 for all the fits we obtain a rate constant 𝑘 that can be compared to the retardation factor −1 (Figs. 7 and 8 in [118]). We see that the general course is the same but 𝑘 tends to be larger than −1 by a factor up to 5.

Appendix A

105 (a)

(b)

Fig. SA6. (a) Fit with Function #1. (b) comparison of rate constant 𝑘 obtained by fitting with constant Avrami index 𝑛 = 1 and retardation factor obtained in [118].

Fig. SA7 shows that function #2 fits the NSA curves quite well. However, the parameters obtained do not vary in a continuous way, indicating over-determination of the function (too many parameters).

106

Fig. SA7. Fitting using Function #2.

The same applies to a double-stage JMAK function, i.e. a generalisation of Function #1 with two 𝑘 and 𝑛 parameters and a weight factor for the two contributions, leading to 7 free parameters. This allows for a good fit but not to derive meaningful parameters (fit not shown).

A6. Modelling of NSA clustering consider excess vacancy annihilation and vacancy trapping by NSA clusters

In the main text of the current work (Sec. 4.2.2.2), we showed that NSA hardening is delayed after applying PA, and especially, to different extents when PA temperature varies while keeping PA hardness constant (as long as >55 HBW). NSA kinetics is faster after PA at higher temperature (for a shorter time) than at lower temperature (for a longer time), with a factor of

~3 for PA at 160 °C and 80 °C, whereas the equilibrium mobile vacancy site fraction at 160 °C is ~58 times higher than at 80 °C.

Here we discuss whether such discrepancy can be explained by vacancy losses due to annihilation at sinks or vacancy trapping in NSA clusters (Zurob’s model [93]). The kinetic master equation (Eq. (4.1) in the main text) is simplified here to Eq. (SA1) using 𝛼𝑁𝑆𝐴 to

Appendix A

107 represent the relative fraction of total clustered solutes during NSA, and (1 − 𝛼𝑁𝑆𝐴) as the simple function related to supersaturation. i.e. 𝜑 ≡ 1. Since we have a non-equilibrium excess vacancy site fraction, which is influenced concurrently by annihilation kinetics and trapping, we solve the master equation numerically instead of analytically.

𝑑𝛼𝑁𝑆𝐴

𝑑𝑡 = 𝑎 × (1 − 𝛼𝑁𝑆𝐴) × 𝑐vac(𝑡). (SA1) Vacancy annihilation during NSA

Two excess vacancy annihilation models were found in the literature, proposed by Schulze et al. [87, 88] and Fischer et al. [78] respectively. Schulze’s model assumes that the annihilation rate of vacancies is proportional to the excess vacancy site fraction to a power of 𝑞 (Eq. (SA2), where 𝑞 = 1 is the most simple case [87]), while Fischer’s model gives an annihilation rate in Eq. (SA3) if we consider only annihilation at dislocation jogs and no hydrostatic stress. 𝑏1, 𝑏2 are both constants governing kinetics. For both models, the difference between the vacancy site fractions after PA at 160 °C and 80 °C starts with ~58 at the beginning of NSA and gradually goes to 1 during NSA, as illustrated in Fig. SA8. The difference between the two models is Fischer’s model eliminates the vacancy discrepancy faster than Schulze’s model (𝑞 = 1). Thus, Fischer’s model will be used in the following discussion.

𝑑𝑐vac(𝑡)

𝑑𝑡 = 𝑏1× (𝑐vac(𝑡) − 𝑐vaceq)𝑞 (SA2)

𝑑𝑐vac(𝑡)

𝑑𝑡 = 𝑏2× 𝑐vac(𝑡) × ln (𝑐vac(𝑡)

𝑐vaceq ) (SA3)

(a) (b)

Fig. SA8. (a) Schematic evolution of vacancy fraction during NSA after PA at different temperatures.

𝑏1 and 𝑏2 are adjusted to produce similar annihilation kinetics after PA at 80 °C by both models.

Fischer’s model gives much faster annihilation after PA at 160 °C. (b) Ratio of the vacancy site fractions during NSA after PA at different temperatures.

108

Vacancy trapping by NSA clusters

Here we take into account that vacancies are trapped by emerging NSA clusters. One model describing such phenomenon was established by Zurob et al. [93] by assuming that the probability of a trapped vacancy to escape from a cluster is proportional to the Boltzmann weight of the interaction energy between the cluster and the vacancy, which is proportional to the number of solute atoms in the cluster due to the number of successful jumps required to escape from the cluster. If the number of clusters remains constant and just their size increases during NSA, then we can express the mobile vacancy site fraction by Eq. (SA4), where 𝑚 and 𝑛 are adjustable constants, and 𝑛 is negative. Note that this model is based on kinetic arguments, which is different from the thermodynamic approach proposed by Pogatscher et al. [83], although they result in similar equations. Since only mobile vacancies can annihilate, 𝑐vac(𝑡) in Eq. (SA3) has to be replaced by 𝑐vac,mob(𝑡) when vacancy trapping is considered.

𝑐vac,mob(𝑡)

𝑐vac(𝑡) = 𝑚 × exp(𝑛 × 𝛼𝑁𝑆𝐴). (SA4) Using the above mentioned models, i.e. Eq. (SA3) and (SA4) with 𝑐vac(𝑡) in Eq. (SA3) replaced by 𝑐vac,mob(𝑡), Fig. SA9 demonstrates the simulated kinetics during NSA after PA at two representative temperatures (160 °C and 80 °C) under various vacancy annihilation and trapping effects by changing the adjustable constant parameters. The initial kinetic difference stems from different equilibrium vacancy site fractions at 160 °C and 80 °C (factor 58). This difference remains (weak annihilation or strong trapping) or gets larger, but is not reduced to 3 as obtained from the experiments.

Appendix A

109 Fig. SA9. Schematic representation of 𝛼𝑁𝑆𝐴 as a function of 𝑡𝑁𝑆𝐴 after PA at 160 °C and 80 °C under various vacancy annihilation and trapping conditions. Stronger trapping: 𝑛 decreases (more negative).

Faster annihilation: 𝑏2 increases.

110

Appendix B

Supplemental material regarding effect of quench rate on multi-stage ageing

B1. Additional TEM images and EDX analysis

Appendix B

111 Fig. SB1. (a – d) Bright-field TEM images of samples after (a) AC, (b) AC and AA for 4 h, (c) IWQ, (d) IWQ and AA for 4 h. (e – f) EDX spectrum of the particles shown in (a).

Two tpyes of particles (P1 and P2) can be observed in the as-quenched state AC sample but only onpe type (P1) can be seen in IWQ sample (Fig. SB1a and c). EDX analysis (Fig. SB1e and f) show that P1 contains Si, Mn, and Fe, which is typical intermetallic particles (dispersoids) in 6xxx alloys. P2 contains mainly Mg and Si, which should be quenched-in precipitates heterogeneously grown on dispersoids. After AA for 4 h (Fig. SB1b and d), needle precipitates can be seen in the matrix. Around dispersoids there are precipitate free zones. The width of these zones is seen larger if quench is slower.

B2. PALS measurements of T2-type quenched samples during NA

Fig. SB2 shows how positron lifetimes evolve at ‘room temperature’ after slow quenches (VC and AC) from 540 °C that were interrupted at a given temperatures (100 °C to 400 °C) by a fast IWQ quench to preserve the configuration. Most curves feature an initially low value of 𝜏1c, a subsequent increase and a merger into a curve that is almost the same for all the experiments including those in Fig. 6.7a. The insets show the data used for the extrapolation to zero NA time on a linear time scale. Fig. 6.8b shows the initial and these extrapolated values.

The mechanism of NSA is not discussed in this work because of the lack of a consistent theory.

112

(a)

(b)

Fig. SB2. Positron lifetimes 𝜏1𝑐 measured during NA of samples quenched in a T2 fashion as shown in Fig. 3.3 for (a) VC, and (b) AC.

Appendix B

113 B3. Vacancy site fraction calculation

Calculating vacancy fractions from positron lifetime data usually requires decompositions of lifetime spectra into 2 or more components. In this work, we measure 𝜏1c only, which, however, is very close to the average positron lifetime that one would calculate from individual components:

𝜏1C ≈ 𝜏̅ = 𝐼0𝜏0+ 𝐼𝑣𝑎𝑐𝜏𝑣𝑎𝑐, 𝐼0+ 𝐼𝑣𝑎𝑐 = 1, (SB1) where the index 0 refers to the reduced bulk lifetime and ′𝑣𝑎𝑐′ refers to annihilation in vacancy-related defects. 𝜏𝑣𝑎𝑐= 245 ps assumed [149].

By applying an equation of the two-state trapping model [150] we obtain:

𝜏0 = 1 1

𝜏𝐵 + 𝜅𝑣𝑎𝑐 (SB2)

𝐼𝑣𝑎𝑐 = 1 𝜅𝑣𝑎𝑐 𝜏𝐵 1

𝜏𝑣𝑎𝑐 + 𝜅𝑣𝑎𝑐, (SB3)

where 𝜏𝐵 ≈160 ps is the lifetime in defect-free aluminium, and 𝜅𝑣𝑎𝑐 is the positron trapping rate of the vacancy-related defect. A relationship applies between 𝜅𝑣𝑎𝑐 and the site fraction of the vacancies 𝑥𝑣𝑎𝑐:

𝑥𝑣𝑎𝑐= 𝜅𝑣𝑎𝑐

𝜇 , (SB4)

where 𝜇 is the positron trapping coefficient of a vacancy in aluminium. Combining Eq.(SB1) – (SB4) eventually leads us to the site fraction of vacancies:

𝑥𝑣𝑎𝑐= 𝜏𝐵−𝜏̅

𝜇𝜏𝐵(𝜏̅−𝜏𝑣𝑎𝑐). (SB5)

Using a commonly accepted value for the specific trapping rate in aluminium,  = 250 ps-1 [151], we can calculate the vacancy site fraction from the one-component positron lifetime.

Fig. SB3 shows the results, which, however, should not been taken too literally due to the approximations used (mainly the absence of clusters and usage of 𝜏1C ≈ 𝜏̅). The presence of clusters would contribute with a lifetime component around 215 ps and would require fewer vacancies to explain a given 𝜏̅.

114

Fig. SB3. Vacancy site fraction calculated using Eqs. (SB1-SB4) as a function of positron lifetime assuming validity of the positron trapping model for one trap related to vacancies characterised by a lifetime of 245 ps. The upper three arrows correspond to 𝜏1𝑐 (extrapolated to zero NSA) as measured after three quenches, the lower one marks the equilibrium vacancy fraction at 180 °C as calculated from the formation enthalpy and vibrational entropy of mono-vacancies in Al [114].

B4. Positron lifetimes of Al-Mg alloy after T1-type quenches

Fig. SB4 displays the positron lifetime evolution in a binary Al-Mg alloy after quenching. It is the analogue to the measurements in Fig. 6.7. Obviously, very little evolution of 𝜏1c is observed.

Appendix B

115 Fig. SB4. Positron lifetimes 𝜏1𝑐 during NA of binary Al-0.5Mg (wt.%) alloy after various T1-type quenches.

B5. AA hardening after NA

Alloy 6014 shows the well-known ‘negative effect’ of NA on AA as AA directly after solutionising and quenching leads to a much faster and eventually higher degree of hardening as AA after 1 week of intermediary NA, see Fig. SB5.

Fig. SB5. Hardening curves during AA directly after IWQ and after NA for 1 week.

116

Appendix C

Supplemental material to linear heating experiments

C1. Normalisation of electrical resistivity

Electrical resistivity calculated using Eq. (3.1) assumes the thickness of the sample ℎ = 0.3 mm, the width of the meander 𝑏 = 1 mm, the length of the measured part 𝐿 = 349 mm, and the cross section 𝐴 = ℎ × 𝑏. All these quantities are nominal values as manufactured by cold rolling and laser cutting, and do not account for manufacturing errors and differences in individual samples. This error can be minimised by more precise measurements of the geometries for each sample, or alternatively without measuring the geometry but assuming that the resistivity measured in Sec.4.1.2 (3911 nΩ·cm at 20 °C) applies to all samples. The resistivity at any temperature can be represented as Eq. (SC1)

𝜌 = 𝑈

𝑈20°𝐶× 3911 𝑛𝛺 · 𝑐𝑚 , (SC1) where 𝑈20°C is the voltage of the sample at 20 °C. If room temperature before start is higher than 20 °C, Eq. (SC2) was used to extrapolate the curve to 20 °C.

𝑈20°𝐶 = 𝑈0 + 𝑈′(20 − 𝑇0) , (SC2) with 𝑈0 being the voltage before heating up and 𝑇0 the corresponding temperature. 𝑈 is the temperature coefficient of the voltage which can be obtained by linear fitting of the voltage during the second heating up to 300 °C.

Fig. SC1 shows resistivity during the second ramp obtained without (Fig. SC1a) and with (Fig. SC1b) such normalisation. It is clearly shown that by employing this method differences between various samples are minimised.

Appendix C

117 Fig. SC1. Resistivity change during the second ramp LH at various heating rates calculated (a) by Eq. (3.1) and nominal geometries of the sample, or (b) by fixing the resistivity at 20 °C and apply Eqs. (SC1, SC2).

C2. PALS lifetime offsets caused by different source corrections

Positron lifetimes during NA in AQ samples were measured many times in this work during 3 years. The lifetimes measured in Fig. 7.5 is observed to be lower than those presented before in Fig. 4.6 and Fig. 6.7. All these curves are now replotted together in Fig. SC2 for a better visuallisation. It is seen that the difference between the curves is a systematic offset. It is speculated that this offset is caused by different source corrections as examplified in Ref. [63]

or spectrometer offsets caused by variations of calibrations that are renewed regularly.

Although all source corrections were aimed to yield ~160 ps for pure Al, small differences might still occur. As all lifetime analyses in Sec. 7.1.4 were conducted using the same source correction, this does not change the relative positions of the curves on the figures, and thus also the discussions based on the lifetime evolution, just the absolute lifetime values might be slightly different to other experiments batches. Thus, comparison of lifetimes between different experiments with different source corrections should be done with care.

118

Fig. SC2. Single-component lifetimes during NA in three AQ samples appeared in this thesis. SC1 corresponds to the curve in Fig. 7.5. SC2 corresponds to the curve in Fig. 6.7a, and SC3 corresponds to the curve in Fig. 4.6.

C3. DSC measurements starting from different temperatures

In the main text, DSC curves and the integrals of samples heated from 0 °C and even lower were used to compare the properties obtained by other measurements such as resistivity and hardness.

Here we show in Fig. SC3 the DSC traces of two AQ samples started from 0 °C and 20 °C respectively. Heating rate was chosen to be 3 K/min because the time spent between 0 °C and 20 °C is the longest. It shows that the starting temeprature has very limited influence on the DSC curve, especially below 300 °C where the features are of most interest. The small differences can be caused different samples or baseline correction, which can be ignored.

Moreover, the instability of DSC device apparently has influenced the first peak in the measurement with starting temperature 20 °C.

Appendix C

119 Fig. SC3. DSC measurements of the AQ samples with different starting temperatures.

C4. Smoothing of resistivity curve for 50 K/min

The resistivity change as a function of temperature during linear heating at 50 K/min is much noisier than other two heating rates. The reason for that is the temperature control of our heating device at high heating rate is less precise than other two. In order to minimise the noise, Savitzky-Golay smoothing using 2 order polynomial function with 10 points in each bin was applied, as demonstrated in Fig. SC4. Smoothed curve has a much smaller variation and has kept the main features of the curve.

Fig. SC4. Smoothing of the resistivity change curve using Savitzky-Golay method with 2 order polynomial function and 10 points per fitting bin.

120

References

[1] http://climate.nasa.gov/vital-signs/carbon-dioxide/, Carbon Dioxide. 2019).

[2] Aluminium in Cars: Unlocking the light-weighting potential, European Aluminium Association, 2012.

[3] W.S. Miller, L. Zhuang, J. Bottema, A.J. Wittebrood, P. De Smet, A. Haszler, A. Vieregge, Recent development in aluminium alloys for the automotive industry, Materials Science and Engineering A 280 (2000) 37-49.

[4] J. Hirsch, Recent developments in aluminium for automotive applications, Transactions of Nonferrous Metals Society of China 24 (2014) 1995-2002.

[5] D. Hanson, M.L.V. Gayler, The constitution and age-hardening of the alloys of aluminium with magnesium and silicon, Journal of Institute of Metals 26(321-359) (1921).

[6] J. Banhart, C.S.T. Chang, Z.Q. Liang, N. Wanderka, M.D.H. Lay, A.J. Hill, Natural Aging in Al-Mg-Si Alloys - A Process of Unexpected Complexity, Advanced Engineering Materials 12(7) (2010) 559-571.

[7] L. Zhen, S.B. Kang, The effect of pre-aging on microstructure and tensile properties of Al-Mg-Si alloys, Scripta Materialia 36(10) (1997) 1089-1094.

[8] L. Zhen, S.B. Kang, H.W. Kim, Effect of natural aging and preaging on subsequent precipitation process of an Al-Mg-Si alloy with high excess silicon, Materials Science and Technology 13(11) (1997) 905-910.

[9] J.D. Bryant, The effects of preaging treatments on aging kinetics and mechanical properties in AA6111 aluminum autobody sheet, Metallurgical and Materials Transactions A 30(8) (1999) 1999-2006.

[10] Y. Birol, Pre-aging to improve bake hardening in a twin-roll cast Al-Mg-Si alloy, Materials Science and Engineering A 391(1-2) (2005) 175-180.

[11] S. Kleiner, C. Henkel, P. Schulz, P.J. Uggowitzer, Paint bake response of aluminium alloy 6016, Aluminium 77(3) (2001) 185-189.

[12] Y. Takaki, T. Masuda, E. Kobayashi, T. Sato, Effects of Natural Aging on Bake Hardening Behavior of Al-Mg-Si Alloys with Multi-Step Aging Process, Materials Transactions 55(8) (2014) 1257-1265.

[13] D.W. Pashley, J.W. Rhodes, A. Sendorek, Delayed ageing in aluminium-magnesium-silicon alloys:

effect on structure and mechnical properties, Journal of the Institute of Metals London 94 (1966) 41-49.

[14] M.J. Starink, Analysis of aluminium based alloys by calorimetry: quantitative analysis of reactions and reaction kinetics, International Materials Reviews 49(3-4) (2004) 191-226.

[15] I. Dutta, S.M. Allen, A Calorimetric Study of Precipitation in Commercial Aluminum Alloy-6061, Journal of Materials Science Letters 10(6) (1991) 323-326.

[16] A.K. Gupta, D.J. Lloyd, Study of precipitation kinetics in a super purity Al-0.8 Pct Mg-0.9 Pct Si alloy using differential scanning calorimetry, Metallurgical and Materials Transactions A 30(3) (1999) 879-884.

[17] D.J. Lloyd, D.R. Evans, A.K. Gupta, Precipitation reactions and the differential scanning calorimetry response of AA6111 alloy, Canadian Metallurgical Quarterly 39(4) (2000) 475-481.

[18] W.F. Miao, D.E. Laughlin, A differential scanning calorimetry study of aluminum alloy 6111 with different pre-aging treatments, Journal of Materials Science Letters 19(3) (2000) 201-203.

[19] S. Esmaeili, D.J. Lloyd, Characterization of the evolution of the volume fraction of precipitates in aged AlMgSiCu alloys using DSC technique, Materials Characterization 55(4-5) (2005) 307-319.

References

121 [20] C.S.T. Chang, J. Banhart, Low-Temperature Differential Scanning Calorimetry of an Al-Mg-Si Alloy, Metallurgical and Materials Transactions A 42A(7) (2011) 1960-1964.

[21] C.S.T. Chang, Z.Q. Liang, E. Schmidt, J. Banhart, Influence of Mg/Si ratio on the clustering kinetics in Al-Mg-Si alloys, International Journal of Materials Research 103(8) (2012) 955-961.

[22] S. Pogatscher, H. Antrekowitsch, H. Leitner, D. Poschmann, Z.L. Zhang, P.J. Uggowitzer, Influence of interrupted quenching on artificial aging of Al-Mg-Si alloys, Acta Materialia 60(11) (2012) 4496-4505.

[23] A. Wilm, Physikalisch-metallurgische Untersuchungen über magnesiumhaltige Aluminiumlegierungen, Metallurgie 8(8) (1911) 225-227.

[24] A. Kussmann, Dritter Teil: Wärmbehandlung, Praktische Metallkunde: Schmelzen und Gießen, spanlose Formung, Wärmbehandlung, Julius Springer, Berlin, 1935.

[25] W. Xu, N. Birbilis, G. Sha, Y. Wang, J.E. Daniels, Y. Xiao, M. Ferry, A high-specific-strength and corrosion-resistant magnesium alloy, Nature Materials 14 (2015) 1229-1235.

[26] S. Jiang, H. Wang, Y. Wu, X. Liu, H. Chen, M. Yao, B. Gault, D. Ponge, D. Raabe, A. Hirata, M.

Chen, Y. Wang, Z. Lu, Ultrastrong steel via minimal lattice misfit and high-density nanoprecipitation, Nature (2017) 460-464.

[27] J.W. Cahn, A Historical Perspective on the Utilization of Phase Diagrams for Precipitation Hardening, Bulletin of Alloy Phase Diagrams 4 (1983) 349-351.

[28] P.D. Merica, R.G. Waltenberg, J.R. Freeman, Constitution and metallography of aluminum and its light alloys with copper and with magnesium, Scientific Papers of the Bureau of Standards 15 (1919) 105-119.

[29] P.D. Merica, R.G. Waltenberg, H. Scott, Heat treatment of Duralumin, Scientific Papers of the Bureau of Standards 15 (1919) 271-316.

[30] A. Guinier, Structure of Age-Hardened Aluminium-Copper Alloys, Nature 142 (1938) 569-570.

[31] G.D. Preston, Nature 142 (1938) 570.

[32] T. Sato, A. Kamio, High resolution electron microscopy of phase decomposition microstructures in aluminium-based alloys, Materials Science and Engineering A 146 (1991) 161-180.

[33] G. Gottstein, Physical Foundations of Materials Science, Springer-Verlag Berlin Heidelberg, 2004.

[34] S. Esmaeili, D.J. Lloyd, W.J. Poole, A yield strength model for the Al-Mg-Si-Cu alloy AA6111, Acta Materialia 51(8) (2003) 2243-2257.

[35] H.W. Zandbergen, S.J. Andersen, J. Jansen, Structure determination of Mg5Si6 particles in Al by dynamic electron diffraction studies, Science 277(5330) (1997) 1221-1225.

[36] J.H. Chen, E. Costan, M.A. van Huis, Q. Xu, H.W. Zandbergen, Atomic pillar-based nanoprecipitates strengthen AlMgSi alloys, Science 312(5772) (2006) 416-419.

[37] A.H. Geisler, J.K. Hill, Analyses and Interpretations of X-ray Diffraction Effects in Patterns of Aged Alloys, Acta Crystallography 1 (1948) 238-252.

[38] G. Thomas, The Ageing Characteristics of Aluminium Alloys - Electron-Transmission Studies of Al-Mg-Si Alloys, Journal of the Institute of Metals 90(2) (1961) 57-63.

[39] G.A. Edwards, K. Stiller, G.L. Dunlop, M.J. Couper, The precipitation sequence in Al-Mg-Si alloys, Acta Materialia 46(11) (1998) 3893-3904.

[40] S.J. Andersen, H.W. Zandbergen, J. Jansen, C. Traeholt, U. Tundal, O. Reiso, The crystal structure of the beta '' phase in Al-Mg-Si alloys, Acta Materialia 46(9) (1998) 3283-3298.

[41] C.D. Marioara, S.J. Andersen, J. Jansen, H.W. Zandbergen, Atomic model for GP-zones in a 6082 Al-Mg-Si system, Acta Materialia 49(2) (2001) 321-328.

[42] S.J. Andersen, C.D. Marioara, A. Froseth, R. Vissers, H.W. Zandbergen, Crystal structure of the orthorhombic U2-Al4Mg4Si4 precipitate in the Al-Mg-Si alloy system and its relation to the beta ' and beta '' phases, Materials Science and Engineering A 390(1-2) (2005) 127-138.

122

[43] M.A. van Huis, J.H. Chen, H.W. Zandbergen, M.H.F. Sluiter, Phase stability and structural relations of nanometer-sized, matrix-embedded precipitate phases in Al-Mg-Si alloys in the late stages of evolution, Acta Materialia 54 (2006) 2945-2955.

[44] M.A. van Huis, J.H. Chen, M.H.F. Sluiter, H.W. Zandbergen, Phase stability and structural features of matrix-embedded hardening precipitates in Al-Mg-Si alloys in the early stages of evolution, Acta Materialia 55(6) (2007) 2183-2199.

[45] R. Vissers, M.A. van Huis, J. Jansen, H.W. Zandbergen, C.D. Marioara, S.J. Andersen, The crystal structure of the beta ' phase in Al-Mg-Si alloys, Acta Materialia 55(11) (2007) 3815-3823.

[46] S.J. Andersen, C.D. Marioara, R. Vissers, M. Torsaeter, R. Bjorge, F.J.H. Ehlers, R. Holmestad, The Dual Nature of Precipitates in Al-Mg-Si Alloys, Materials Science Forum 638-642 (2010) 390-395.

[47] C.D. Marioara, S.J. Andersen, J. Friis, O. Engler, Y. Aruga, The nature of solute clusters and GP-zones in the Al-Mg-Si system, Proceedings of the 16th International Aluminum Alloys Conference (ICAA16) (2018).

[48] C.H. Liu, Y.X. Lai, J.H. Chen, G.H. Tao, L.M. Liu, P.P. Ma, C.L. Wu, Natural-aging-induced reversal of the precipitation pathways in an Al–Mg–Si alloy, Scripta Materialia 115 (2016) 150-154.

[49] Y.X. Lai, B.C. Jiang, C.H. Liu, Z.K. Chen, C.L. Wu, J.H. Chen, Low-alloy-correlated reversal of the precipitation sequence in Al-Mg-Si alloys, Journal of Alloys and Compounds 701 (2017) 94-98.

[50] V. Fallah, A. Korinek, N. Ofori-Opoku, B. Raeisinia, M. Gallerneault, N. Provotas, S. Esmaeili, Atomic-scale pathway of early-stage precipitation in Al-Mg-Si alloys, Acta Materialia 82 (2015) 457-467.

[51] A.I. Morley, M.W. Zandbergen, A. Cerezo, G.D.W. Smith, The effect of pre-ageing and addition of copper on the precipitation behaviour in Al-Mg-Si alloys, Materials Science Forum 519-521 (2006) 543-548.

[52] A. Serizawa, S. Hirosawa, T. Sato, Three-Dimensional Atom Probe Characterization of Nanoclusters Responsible for Multistep Aging Behavior of an Al-Mg-Si Alloy, Metallurgical and Materials Transactions A 39A (2008) 245-251.

[53] M.W. Zandbergen, Q. Xu, A. Cerezo, G.D.W. Smith, Study of precipitation in Al-Mg-Si alloys by atom probe tomography I. Microstructural changes as a function of ageing temperature, Acta Materialia 101 (2015) 136-148.

[54] M.W. Zandbergen, Q. Xu, A. Cerezo, G.D.W. Smith, Data Analysis and Other Considerations Concerning the Study of Precipitation in Al-Mg-Si Alloys by Atom Probe Tomography, Data in Brief 5 (2015) 626-641.

[55] Y. Aruga, M. Kozuka, Y. Takaki, T. Sato, Formation and reversion of clusters during natural aging and subsequent artificial aging in an Al–Mg–Si alloy, Materials Science and Engineering: A 631 (2015) 86-96.

[56] Y. Aruga, S.N. Kim, M. Kozuka, E. Kobayashi, T. Saito, Effects of cluster characteristics on two-step aging behavior in Al-Mg-Si alloys with different Mg/Si ratios and natural aging periods, Materials Science and Engineering A 718 (2018) 371-376.

[57] Y. Aruga, M. Kozuka, T. Sato, Formulation of initial artificial age-hardening response in an Al-Mg-Si alloy based on the cluster classification using a high-detection-efficiency atom probe, Journal of Alloys and Compounds 739 (2018) 1115-1123.

[58] A. Poznak, R.K.W. Marceau, P.G. Sanders, Composition dependent thermal stability and evolution of solute clusters in Al-Mg-Si analyzed using atom probe tomography, Materials Science and Engineering A 721 (2018) 47-60.

[59] S. Esmaeili, D. Vaumousse, M.W. Zandbergen, W.J. Poole, A. Cerezo, D.J. Lloyd, A study on the early-stage decomposition in the Al-Mg-Si-Cu alloy AA6111 by electrical resistivity and three-dimensional atom probe, Philosophical Magazine 87(25-27) (2007) 3797-3816.

References

123 [60] M. Torsaeter, H.S. Hasting, W. Lefebvre, C.D. Marioara, J.C. Walmsley, S.J. Andersen, R.

Holmestad, The influence of composition and natural aging on clustering during preaging in Al-Mg-Si alloys, Journal of Applied Physics 108(7) (2010) 073527.

[61] Y. Aruga, M. Kozuka, Y. Takaki, T. Sato, Effects of natural aging after pre-aging on clustering and bake-hardening behavior in an Al-Mg-Si alloy, Scripta Materialia 116 (2016) 82-86.

[62] Y. Aruga, M. Kozuka, Y. Takaki, T. Sato, Evaluation of Solute Clusters Associated with Bake-Hardening Response in Isothermal Aged Al-Mg-Si Alloys Using a Three-Dimensional Atom Probe, Metallurgical and Materials Transactions A 45A(13) (2014) 5906-5913.

[63] J. Banhart, M.D.H. Lay, C.S.T. Chang, A.J. Hill, Kinetics of natural aging in Al-Mg-Si alloys studied by positron annihilation lifetime spectroscopy, Physical Review B 83(1) (2011) 014101.

[64] M. Liu, J. Cizek, C.S.T. Chang, J. Banhart, Early stages of solute clustering in an Al–Mg–Si alloy, Acta Materialia 91 (2015) 355–364.

[65] C. Panseri, T. Federighi, A Resistometric Study of Preprecipitation in an Aluminium-1.4 Percent Mg2si Alloy, Journal of the Institute of Metals London 94 (1966) 99-197.

[66] C.D. Marioara, S.J. Andersen, J. Jansen, H.W. Zandbergen, The influence of temperature and storage time at RT on nucleation of the beta '' phase in a 6082 Al-Mg-Si alloy, Acta Materialia 51(3) (2003) 789-796.

[67] M.L.V. Gayler, G.D. Preston, The age-hardening of some aluminium alloys, Journal of the Institute of Metals (London) 41 (1929) 191-247

[68] P. Brenner, H. Kostron, Über die Vergütung der Aluminium-Magnesium-Silizium Legierungen (Pantal), Zeitschrift für Metallkunde 31(4) (1939) 89-97.

[69] D.W. Pashley, M.H. Jacobs, J.T. Vietz, The basic processes affecting two-step ageing in an Al-Mg-Si alloy, Philosophical Magazine 16 (1967) 139.

[70] S. Pogatscher, H. Antrekowitsch, H. Leitner, A.S. Sologubenko, P.J. Uggowitzer, Influence of the thermal route on the peak-aged microstructures in an Al-Mg-Si aluminum alloys, Scripta Materialia 68 (2013) 158-161.

[71] M. Madanat, M. Liu, J. Banhart, Reversion of natural ageing in Al-Mg-Si alloys, Acta Materialia 159 (2018) 163-172.

[72] S. Pogatscher, H. Antrekowitsch, H. Leitner, T. Ebner, P.J. Uggowitzer, Mechanisms controlling the artificial aging of Al-Mg-Si Alloys, Acta Materialia 59(9) (2011) 3352-3363.

[73] Y. Yan, Investigation of the negative and positive effects of natural ageing on artificial ageing response in Al-Mg-Si alloys, Technische Universität Berlin, Berlin, 2014.

[74] L. Ding, Y. Weng, S. Wu, R.E. Sanders, Z. Jia, Q. Liu, Influence of interrupted quenching and pre-ageing on the bake hardening of Al-Mg-Si alloy, Materials Science and Engineering A 651 (2016) 991-998.

[75] M. Murayama, K. Hono, Pre-precipitate clusters and precipitation processes in Al-Mg-Si alloys, Acta Materialia 47(5) (1999) 1537-1548.

[76] Y. Aruga, M. Kozuka, Y. Takaki, T. Sato, Formation and reversion of clusters during natural aging and subsequent artificial aging in an Al-Mg-Si alloy, Materials Science and Engineering A 631 (2015) 86-96.

[77] C.S.T. Chang, I. Wieler, N. Wanderka, J. Banhart, Positive effect of natural pre-ageing on precipitation hardening in Al-0.44 at% Mg-0.38 at% Si alloy, Ultramicroscopy 109(5) (2009) 585-592.

[78] F.D. Fischer, J. Svoboda, F. Appel, E. Kozeschnik, Modeling of excess vacancy annihilation at different types of sinks, Acta Materialia 59 (2011) 3463-3472.

[79] W. Desorbo, H.N. Treaftis, D. Turnbull, Rate of Clustering in Al-Cu Alloys at Low Temperatures, Acta Metallurgica 6(6) (1958) 401-413.

[80] T. Federighi, Quenched-in vacancies and rate of formation of zones in aluminum alloys, Acta Metallurgica 6 (1958) 379-381.

124

[81] A. Kelly, R.B. Nicholson, Precipitation hardening, Progress in Materials Science 10 (1963) 149-391.

[82] S. Pogatscher, E. Kozeschnik, H. Antrekowitsch, M. Werinos, S.S.A. Gerstl, J.F. Loffler, P.J.

Uggowitzer, Process-controlled suppression of natural aging in an Al-Mg-Si alloy, Scripta Materialia 89 (2014) 53-56.

[83] S. Pogatscher, H. Antrekowitsch, M. Werinos, F. Moszner, S.S.A. Gerstl, M.F. Francis, W.A.

Curtin, J.F. Löffler, P.G. Uggowitzer, Diffusion on Demand to Control Precipitation Aging: Application to Al-Mg-Si Alloys, Physical Review Letters 112 (2014) 225701.

[84] M. Werinos, H. Antrekowitsch, T. Ebner, R. Prillhofer, W.A. Curtin, P.J. Uggowitzer, S.

Pogatscher, Design strategy for controlled natural aging in Al-Mg-Si alloys, Acta Materialia 118 (2016) 296-305.

[85] M. Liu, X. Zhang, B. Körner, M. Elsayed, Z. Liang, D. Leyvraz, J. Banhart, Effect of Sn and In on the natural ageing kinetics of Al-Mg-Si alloys, Materialia 6 (2019) 100261.

[86] M. Werinos, H. Antrekowitsch, T. Ebner, R. Prillhofer, P.J. Uggowitzer, S. Pogatscher, Hardening of Al-Mg-Si alloys: Effect of trace elements and prolonged natural ageing, Materials and Design 107 (2016) 257-268.

[87] H.A. Schulze, K. Lücke, Short-Range-Order Formation in Dilute Alloys due to Quenched-in Vacancies, Journal of Applied Physics 39 (1968) 4860-4862.

[88] A. Schulze, K. Lücke, The influence of vacancies on short-range order formation in Au-Ag alloys, Acta Metallurgica 20 (1972) 529-542.

[89] P.B. Hirsch, J. Silcox, R.E. Smallman, K.H. Westmacott, Dislocation loops in quenched aluminium, Philosophical Magazine 8 (1958) 897-908.

[90] G. Thomas, Quenching defects in binary aluminium alloys, Philosophical Magazine 4 (1959) 1213-1228.

[91] K.H. Westmacott, R.L. Peck, A rationalization of secondary defect structures in aluminium-based alloys, Philosophical Magazine 23 (1971) 611-622.

[92] W. Sun, Y. Zhu, R.K.W. Marceau, L. Wang, Q. Zhang, X. Gao, C. Hutchinson, Precipitation strengthening of aluminium alloys by room-temperature cyclic plasticity, Science 363 (2019) 972-975.

[93] H.S. Zurob, H. Seyedrezai, A model for the growth of solute clusters based on vacancy trapping, Scripta Materialia 61(2) (2009) 141-144.

[94] W.M. Lomer, Point defects and diffusion in metals and alloys, Vacancies and other point defects in metals and alloys: a symposium held at the Atomic Energy Research Establishment, Harwell, Berks., Institute of Metals, London, 1958, pp. 79-98.

[95] C. Wolverton, Solute-vacancy binding in aluminum, Acta Materialia 55 (2007) 5867-5872.

[96] S. Hirosawa, F. Nakamura, T. Sato, First-principles calculation of interaction energies between solutes and/or vacancies for predicting atomistic behaviors of microalloying elements in aluminum alloys, Materials Science Forum 561-565 (2007) 283-286.

[97] I. Erdle, Nicht-isotherme Aushärtung einer 6014 Aluminiumlegierung, Master thesis, TU Berlin (2019).

[98] A. Deschamps, M. Garcia, J. Chevy, B. Davo, F. De Geuser, Influence of Mg and Li content on the microstructure evolution of Al-Cu-Li alloys during long-term ageing, Acta Materialia 122 (2017) 32-46.

[99] R. Ivanov, A. Deschamps, F. De Geuser, Clustering kinetics during natural ageing of Al-Cu based alloys with (Mg, Li) additions, Acta Materialia 157 (2018) 186-195.

[100] S. Pogatscher, H. Antrekowitsch, P.J. Uggowitzer, Influence of starting temperature on differential scanning calorimetry measurements of an Al-Mg-Si alloy, Materials Letters 100 (2013) 163-165.

References

125 [101] M. Liu, Clustering kinetics in Al-Mg-Si alloys investigated by positron annihilation techniques, PhD thesis, Technische Universität Berlin, Berlin, 2014.

[102] M. Saga, Y. Sasaki, M. Kikuchi, Z. Yan, M. Matsuo, Effect of pre-aging temperature on the behavior in the early stage of aging at high temperature for Al-Mg-Si alloy, Materials Science Forum 217 (1996) 821-826.

[103] A. Falahati, P. Lang, E. Kozeschnik, Precipitation in Al-alloy 6016 - the role of excess vacancies, Materials Science Forum 706-709 (2012) 317-322.

[104] I. Kovács, J. Lendvai, E. Nagy, The Mechanism of Clustering in Supersaturated Solid-Solutions of Al-Mg2Si Alloys, Acta Metallurgica 20(7) (1972) 975-983.

[105] M. Mantina, Y. Wang, L.Q. Chen, Z.K. Liu, C. Wolverton, First principles impurity diffusion coefficients, Acta Materialia 57(14) (2009) 4102-4108.

[106] L.F. Cao, P.A. Rometsch, M.C. Couper, Clustering behaviour in an Al-Mg-Si-Cu alloy during natuiral ageing and subsequent under-ageing, Materials Science and Engineering A 559 (2013) 257-261.

[107] S. Kim, J. Kim, H. Tezuka, E. Kobayashi, T. Sato, Formation Behavior of Nanoclusters in Al-Mg-Si Alloys with Different Mg and Si Concentration, Materials Transactions 54(3) (2013) 297-303.

[108] M.J. Starink, On the meaning of the impingement parameter in kinetic equations for nucleation and growth reactions, Journal of Materials Science 36(18) (2001) 4433-4441.

[109] M.J. Starink, L.F. Cao, P.A. Rometsch, A model for the thermodynamics of and strengthening due to co-clusters in Al-Mg-Si-based alloys, Acta Materialia 60(10) (2012) 4194-4207.

[110] M.J. Starink, S.C. Wang, The thermodynamics of and strengthening due to co-clusters: General theory and application to the case of Al-Cu-Mg alloys, Acta Materialia 57 (2009) 2376-2389.

[111] R. Ferragut, A. Somoza, I. Torriani, Pre-precipitation study in the 7012 Al-Zn-Mg-Cu alloy by electrical resistivity, Materials Science and Engineering A 334(1-2) (2002) 1-5.

[112] W. Sha, The use of resistivity data in calculating the kinetics of precipitate evolution in aluminium-copper-magnesium alloys based on Johnson-Mehl-Avrami theory, Physica Status Solidi a-Applications and Materials Science 202(10) (2005) 1903-1908.

[113] A.R. Eivani, A.K. Taheri, Modeling age hardening kinetics of an Al-Mg-Si-Cu aluminum alloy, Journal of Materials Processing Technology 205(1-3) (2008) 388-393.

[114] T. Hehenkamp, Absolute vacancy concentrations in noble metals and some of their alloys, Journal of Physics and Chemistry of Solids 55(10) (1994) 907-915.

[115] A.B. Lidiard, The influence of solutes on self-diffusion in metals, Philosophical Magazine 5(59) (1960) 1171-1180.

[116] T. Weisz, P. Warczok, T. Ebner, A. Falahati, E. Kozeschnik, Simulation of Natural Ageing in Al-Mg-Si alloys, Materials Science Forum 828-829 (2015) 468-473.

[117] M.J. Starink, A.M. Zahra, An analysis method for nucleation and growth controlled reactions at constant heating rate, Thermochimica Acta 292(1-2) (1997) 159-168.

[118] Z. Yang, Z. Liang, D. Leyvraz, J. Banhart, Effect of pre-ageing on natural secondary ageing and paint-bake hardening in Al-Mg-Si alloys, Materialia 7 (2019) 100413.

[119] H. Hirasawa, Precipitation process of Al-Mn and Al-Cr supersaturated solid solution in presence of age-hardening phases, Scripta Metallurgica 9 (1975) 955-958.

[120] L. Lodgaard, N. Ryum, Precipitation of dispersoids containing Mn and/or Cr in Al-Mg-Si alloys, Materials Science and Engineering A 283 (2000) 144-152.

[121] B. Milkereit, N. Wanderka, C. Schick, O. Kessler, Continuous cooling precipitation diagrams of Al-Mg-Si alloys, Materials Science and Engineering A 550 (2012) 87-96.

[122] B. Milkereit, M.J. Starink, Quench sensitivity of Al-Mg-Si alloys: A model for linear cooling and strengthening, Materials & Design 76 (2015) 117-129.

126

[123] K. Strobel, M.A. Easton, M.D.H. Lay, P.A. Rometsch, S. Zhu, L. Sweet, N.C. Parson, A.J. Hill, Quench sensitivity in a dispersoid-containing Al-Mg-Si alloy, Metallurgical and Materials Transaction A 50A (2019) 1957-1969.

[124] A. Dons, O. Lohne, Quench sensitivity of AlMgSi-alloys containing Mn or Cr, MRS Proceedings 21 (1983) 723-728.

[125] S. Zajac, B. Bengtsson, C. Jönsson, Influence of cooling after homogenisation and reheating to extrusion on extrudability and final properties of AA6063 and AA6082 alloys, Materials Science Forum 396-402 (2002) 399-404.

[126] K. Strobel, M.A. Easton, L. Sweet, M.J. Couper, J.F. Nie, Relating quench sensitivity to microstructure in 6000 series aluminium alloys, Materials Transactions 52 (2011) 914-919.

[127] P.A. Rometsch, S.C. Wang, A. Harriss, P.J. Gregson, M.J. Starink, The effect of homogenizing on the quench sensitivity of 6082, Materials Science Forum 396-4 (2002) 655-660.

[128] J. Buha, P.R. Munroe, R.N. Lumley, A.G. Crosky, A.J. Hill, Positron studies of precipitation in 6061 aluminium alloys, in: B.C. Muddle, A.J. Morton, J.-F. Nie (Eds.) International Conference on Aluminium Alloys (ICAA-9), Institute of Materials Engineering Australia, Brisbane, Australia, 2004, pp. 1028-1033.

[129] M.D.H. Lay, H.S. Zurob, C.R. Hutchinson, T.J. Bastow, A.J. Hill, Vacancy Behavior and Solute Cluster Growth During Natural Aging of an Al-Mg-Si Alloy, Metallurgical and Materials Transactions A 43A(12) (2012) 4507-4513.

[130] J. Silcox, M.J. Whelan, Direct observations of the annealing of prismatic dislocation loops and of climb of dislocations in quenched aluminium, Philosophical Magazine 5 (1960) 1-23.

[131] J.M. Dowling, J.W. Martin, The influence of Mn additions on the deformation behaviour of an Al-Mg-Si alloy, Acta Metallurgica 24 (1976) 1147-1153.

[132] K.H. Westmacott, R.S. Barnes, D. Hull, R.E. Smallman, Vacancy trapping in quenched aluminium alloys, Philosophical Magazine 6 (1961) 929-935.

[133] R.M.J. Cotterill, R.L. Segall, The effect of quenching history, quenching temperature and trace impurities on vacancy clusters in aluminium and gold, Philosophical Magazine 8 (1963) 1105-1125.

[134] M. Kiritani, Formation of voids and dislocation loops in quenched aluminum, Journal of the Physical Society of Japan 19 (1964) 618-631.

[135] M. Kiritani, T. Nishikawa, S. Yoshida, Formation of dislocation loops at low temperatures in quenched aluminum, Journal of the Physical Society of Japan 27 (1969) 67-73.

[136] M. Liu, B. Klobes, J. Banhart, Positron lifetime study of the formation of vacancy clusters and dislocations in quenched Al, Al-Mg and Al-Si alloys, Journal of Materials Science 51 (2016) 7754-7767.

[137] Z. Matyas, Change of electrical resistance of alloys during ageing, Philosophical Magazine 40 (1949) 324-337.

[138] H. Herman, J.B. Cohen, Resistivity changes due to formation of G. P. Zones, Nature 191 (1961) 63-64.

[139] H. Herman, M.E. Fine, J.B. Cohen, Formation and Reversion of Guinier-Preston Zones in Al-5.3 At.% Zn, Acta Metallurgica 11(1) (1963) 43-56.

[140] A.K. Gupta, D.J. Lloyd, S.A. Court, Precipitation hardening processes in an Al-0.40%Mg-1.3%Si-0.25%Fe aluminum alloy, Materials Science and Engineering A 301(2) (2001) 140-146.

[141] X. Zhang, M. Liu, H. Sun, J. Banhart, Influence of Sn on the age hardening behavior of Al-Mg-Si alloys at different temperatures, Materialia 8 (2019) 100441.

[142] C.H. Liu, P.P. Ma, L.H. Zhan, M.H. Huang, J.J. Li, Solute Sn-induced formation of composite β’/β’’ precipitates in Al-Mg-Si alloy, Scripta Materialia 155 (2018) 68-72.

[143] C. Haase, H. Wurst, Zur Frage der Kalt- und Warmaushärtung bei Aluminium-Magnesium-Silizium-Legierungen, Zeitschrift für Metallkunde 33(12) (1941) 399-403.

Im Dokument Multi-stage ageing in an Al-Mg-Si alloy (Seite 109-141)