Skip to main content
Log in

Coarsening behavior of precipitate Al3(Sc,Zr) in supersaturated Al-Sc-Zr alloy via melt spinning and extrusion

  • Metals & corrosion
  • Published:
Journal of Materials Science Aims and scope Submit manuscript

Abstract

Al-Sc-Zr alloy contains precipitate Al(Sc,Zr) shows good coarsening resistance due to the small diffusivities of Sc and Zr. Two supersaturated Al-0.4Sc-0.4Zr at% alloys were produced using melt-spinning and following extrusion. The coarsening behavior of supersaturated Al-0.4Sc-0.4Zr at% alloys was investigated using scanning transmission electron microscopy (STEM). Accelerated coarsening was observed in the extruded melt-spun ribbon with larger precipitate size and higher precipitate volume fraction than in melt-spun ribbon. The accelerated effective diffusivity after extrusion was explicitly linked to the initially refined extruded microstructure. Precipitate volume fraction and chemical composition in both melt-spun ribbon and extruded rod during annealing were quantified with the help of STEM and energy-dispersive x-ray spectroscopy. Precipitate size and number density were predicted using Umantsev-Olson-Kuehman-Voorhees model for ternary alloys, and the predictions showed good agreement with experimental results by considering changes in diffusivity and precipitate volume fraction.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13
Figure 14
Figure 15

Similar content being viewed by others

References

  1. Fuller CB, Seidman DN, Dunand DC (2003) Mechanical properties of Al (Sc, Zr) alloys at ambient and elevated temperatures. Acta Mater 51(16):4803–4814

    Article  CAS  Google Scholar 

  2. Knipling KE, Seidman DN, Dunand DC (2011) Ambient- and high- temperature mechanical properties of isochronally aged Al-0.06Sc, Al-0.06 Zr, and Al-0.06Sc-0.06Zr Alloys (at%). Acta Mater 59(3):943–954. https://doi.org/10.1016/j.actamat.2010.10.017

    Article  CAS  Google Scholar 

  3. Novotny GM, Ardell AJ (2001) Precipitation of Al3Sc in binary Al–Sc alloys. Mater Sci Eng, A 318(1–2):144–154. https://doi.org/10.1016/S0921-5093(01)01326-0

    Article  Google Scholar 

  4. Knipling KE, Karnesky RA, Lee CP, Dunand DC, Seidman DN (2010) Precipitation evolution in Al-0.1Sc, Al-0.1Zr, and Al-0.1Sc-0.1Zr (at%) alloys during isochronal aging. Acta Mater 58(15):5184–5195. https://doi.org/10.1016/j.actamat.2010.05.054

    Article  CAS  Google Scholar 

  5. Clouet E, Laé L, Épicier T, Lefebvre W, Nastar M, Deschamps A (2006) Complex precipitation pathways in multicomponent alloys. Nat Mater 5(6):482–488. https://doi.org/10.1038/nmat1652

    Article  CAS  Google Scholar 

  6. Knipling KE, Dunand DC (2008) Creep resistance of cast and aged Al–0.1Zr and Al–0.1Zr–0.1Ti (at.%) alloys at 300–400 °C. Scr Mater 59(4):387–390. https://doi.org/10.1016/j.scriptamat.2008.02.059

    Article  CAS  Google Scholar 

  7. Fujikawa S-I (1996) Solid state diffusion in light metals. I. J Jpn Inst Light Metals (Jpn) 46(4):202–215

    CAS  Google Scholar 

  8. Grammatikakis JG, Eftaxias K, Patapis S, Hadjicontis V (1988) A simple model applied to the elastic properties of the lead-thallium and magnesium-lithium alloys. J Phys Chem Solids 49(8):965–967. https://doi.org/10.1016/0022-3697(88)90014-5

    Article  CAS  Google Scholar 

  9. Hirano K, Fujikawa SI (1978) Impurity diffusion in aluminum. J Nucl Mater 69:564–566. https://doi.org/10.1016/0022-3115(78)90275-1

    Article  Google Scholar 

  10. Fujikawa S, Hirano K, Fukushima Y (1978) Diffusion of silicon in aluminum. MTA 9(12):1811–1815. https://doi.org/10.1007/BF02663412

    Article  Google Scholar 

  11. Fujikawa SI, Hirano K (1991) Impurity-diffusion of copper in aluminum. Defect Diffus Forum 66–69:447

    Article  Google Scholar 

  12. Tolley A, Radmilovic V, Dahmen U (2005) Segregation in Al3(Sc, Zr) precipitates in Al–Sc–Zr alloys. Scr Mater 52(7):621–625. https://doi.org/10.1016/j.scriptamat.2004.11.021

    Article  CAS  Google Scholar 

  13. Voorhees P (2006) Alloys: scandium overtakes zirconium. Nat Mater 5:435–436. https://doi.org/10.1038/nmat1663

    Article  CAS  Google Scholar 

  14. Marquis EA, Seidman DN (2001) Nanoscale structural evolution of Al3Sc precipitates in Al(Sc) alloys. Acta Mater 49(11):1909–1919. https://doi.org/10.1016/S1359-6454(01)00116-1

    Article  CAS  Google Scholar 

  15. Okamoto H (2002) Al-Zr (Aluminum-Zirconium). J Phase Equilib 23:455–456

    Article  CAS  Google Scholar 

  16. Murray JL (1998) The Al-Sc (aluminum-scandium) system. JPE 19(4):380–384. https://doi.org/10.1361/105497198770342120

    Article  CAS  Google Scholar 

  17. Murray J, Peruzzi A, Abriata JP (1992) The Al-Zr (aluminum-zirconium) system. JPE 13(3):277–291. https://doi.org/10.1007/BF02667556

    Article  CAS  Google Scholar 

  18. Yin Z, Pan Q, Zhang Y, Jiang F (2000) Effect of minor Sc and Zr on the microstructure and mechanical properties of Al–Mg based alloys. Mater Sci Eng, A 280(1):151–155. https://doi.org/10.1016/S0921-5093(99)00682-6

    Article  Google Scholar 

  19. Knipling KE, Dunand DC, Seidman DN (2006) Criteria for developing castable, creep-resistant aluminum-based alloys–A review. Zeitschrift für Metallkunde 97(3):246–265

    Article  CAS  Google Scholar 

  20. Aziz MJ (1982) Model for solute redistribution during rapid solidification. J Appl Phys 53(2):1158–1168. https://doi.org/10.1063/1.329867

    Article  CAS  Google Scholar 

  21. Toropova LS (1998) Advanced aluminum alloys containing scandium: structure and properties. Gordon and Breach Science Publishers, London

    Google Scholar 

  22. Yang Y, Licavoli JJ, Sanders PG (2020) Improved strengthening in supersaturated Al-Sc-Zr alloy via melt-spinning and extrusion. J Alloy Compd 826:154185. https://doi.org/10.1016/j.jallcom.2020.154185

    Article  CAS  Google Scholar 

  23. Kawamura Y, Inoue A, Sasamori K, Masumoto T (1993) High-strength powder metallurgy aluminum alloys in glass-forming Al-Ni-Ce-(Ti or Zr) systems. Scr Metall Mater 29(2):275–280

    Article  CAS  Google Scholar 

  24. Szymański W, Bigaj M, Gawlik M, Mitka M, Szymanek M (2014) Consolidation by continuous rotary extrusion of aluminium alloys cast by the melt spinning process. Arch Metall Mater. https://doi.org/10.2478/amm-2014-0050

    Article  Google Scholar 

  25. Zribi Z, Ktari HH, Herbst F, Optasanu V, Njah N (2019) EBSD, XRD and SRS characterization of a casting Al-7wt%Si alloy processed by equal channel angular extrusion: dislocation density evaluation. Mater Charact 153:190–198. https://doi.org/10.1016/j.matchar.2019.04.044

    Article  CAS  Google Scholar 

  26. Berndt N, Frint P, Wagner MF-X (2018) Influence of extrusion temperature on the aging behavior and mechanical properties of an AA6060 aluminum alloy. Metals 8(1):51. https://doi.org/10.3390/met8010051

    Article  CAS  Google Scholar 

  27. Lifshitz IM, Slyozov VV (1961) The kinetics of precipitation from supersaturated solid solutions. J Phys Chem Solids 19(1):35–50. https://doi.org/10.1016/0022-3697(61)90054-3

    Article  Google Scholar 

  28. C. Wagner, “Theorie der Alterung von Niederschlägen durch Umlösen (Ostwald-Reifung),” Zeitschrift für Elektrochemie, Berichte der Bunsengesellschaft für physikalische Chemie, vol. 65, no. 7–8, pp. 581–591, 1961, doi: https://doi.org/10.1002/bbpc.19610650704

  29. Umantsev A, Olson GB (1993) Ostwald ripening in multicomponent alloys. Scr Metall Mater 29(8):1135–1140. https://doi.org/10.1016/0956-716X(93)90191-T

    Article  CAS  Google Scholar 

  30. Kuehmann CJ, Voorhees PW (1996) Ostwald ripening in ternary alloys. MMTA 27(4):937–943. https://doi.org/10.1007/BF02649761

    Article  Google Scholar 

  31. Forbord B, Hallem H, Røyset J, Marthinsen K (2008) Thermal stability of Al3(Scx, Zr1−x)-dispersoids in extruded aluminium alloys. Mater Sci Eng A 475(1):241–248. https://doi.org/10.1016/j.msea.2007.04.054

    Article  CAS  Google Scholar 

  32. Fuller C, Seidman D (2005) Temporal evolution of the nanostructure of Al(Sc, Zr) alloys: Part II-coarsening of Al(ScZr) precipitates. Acta Mater 53:5415–5428. https://doi.org/10.1016/j.actamat.2005.08.015

    Article  CAS  Google Scholar 

  33. Ashby MF, Brown LM (1963) Diffraction contrast from spherically symmetrical coherency strains. Philos Mag A J Theor Exp Appl Phys 8(91):1083–1103. https://doi.org/10.1080/14786436308207338

    Article  Google Scholar 

  34. Ashby MF, Brown LM (1963) On diffraction contrast from inclusions. Philos Mag A J Theor Exp Appl Phys 8(94):1649–1676. https://doi.org/10.1080/14786436308207329

    Article  CAS  Google Scholar 

  35. Yang Y, Mukherjee P, Sanders P (2019) Precipitation hardening via chemically ordered L12 precipitates in Al-Sc-Zr Alloys— new insights using combined STEM and EDS study. Microsc Microanal 25(S2):2188–2189. https://doi.org/10.1017/S143192761901167X

    Article  Google Scholar 

  36. Campbell CE, Bendersky LA, Boettinger WJ, Ivester R (2006) Microstructural characterization of Al-7075-T651 chips and work pieces produced by high-speed machining. Mater Sci Eng, A 430(1):15–26. https://doi.org/10.1016/j.msea.2006.04.122

    Article  CAS  Google Scholar 

  37. Brown AM, Ashby MF (1980) Correlations for diffusion constants. Acta Metall 28(8):1085–1101. https://doi.org/10.1016/0001-6160(80)90092-9

    Article  CAS  Google Scholar 

  38. Balluffi RW (1970) On measurements of self-diffusion rates along dislocations in F.C.C. Metals. Phys Status Solidi (b) 42(1):11–34. https://doi.org/10.1002/pssb.19700420102

    Article  CAS  Google Scholar 

  39. Murch G, Nowick A (1984) Diffusion in crystalline solids. Academic Press Inc, Orlando, Florida, p 482

    Google Scholar 

  40. Footner PK, Richards BP (1982) Long — term growth of superalloy γ′ particles. J Mater Sci 17(7):2141–2153. https://doi.org/10.1007/BF00540433

    Article  CAS  Google Scholar 

  41. R. Wagner, R. Kampmann, P. W. Voorhees, and G. Kostorz, “Homogeneous Second Phase Precipitation: Phase Transformations in Materials Science and Technology,” 2001, Accessed: Feb. 08, 2020. [Online]. Available: https://www.scholars.northwestern.edu/en/publications/homogeneous-second-phase-precipitation-phase-transformations-in-m.

  42. Balanda KP, MacGillivray HL (1988) Kurtosis: a critical review. Am Stat 42(2):111–119. https://doi.org/10.2307/2684482

    Article  Google Scholar 

  43. D. Dugar, “Skew and Kurtosis: 2 Important Statistics terms you need to know in Data Science,” Medium, Jul. 18, 2020. https://codeburst.io/2-important-statistics-terms-you-need-to-know-in-data-science-skewness-and-kurtosis-388fef94eeaa Accessed from 28 Jul 2020.

  44. Ardell AJ (1972) The effect of volume fraction on particle coarsening: theoretical considerations. Acta Metall 20(1):61–71. https://doi.org/10.1016/0001-6160(72)90114-9

    Article  Google Scholar 

  45. H. Okamoto, Phase Diagrams of Dilute Binary Alloys. ASM International, 2002.

  46. Fujikawa SI (1997) Impurity diffusion of scandium in aluminium. Defect Diffus Forum 143–147:115–120

    Article  Google Scholar 

  47. Deschamps A, Lae L, Guyot P (2007) In situ small-angle scattering study of the precipitation kinetics in an Al–Zr–Sc alloy. Acta Mater 55(8):2775–2783. https://doi.org/10.1016/j.actamat.2006.12.015

    Article  CAS  Google Scholar 

  48. Sen PK (1968) Estimates of the regression coefficient based on Kendall’s Tau. J Am Stat Assoc 63(324):1379–1389. https://doi.org/10.1080/01621459.1968.10480934

    Article  Google Scholar 

  49. H. Theil, “A rank-invariant method of linear and polynomial regression analysis. I,” Nederl. Akad. Wetensch., Proc., vol. 53, pp. 386–392 = Indagationes Math. 12, 85–91 (1950), 1950.

  50. Fuller CB, Murray JL, Seidman DN (2005) Temporal evolution of the nanostructure of Al(Sc, Zr) alloys: part I - chemical compositions of Al3(Sc1−xZrx) precipitates. Acta Mater 53(20):5401–5413. https://doi.org/10.1016/j.actamat.2005.08.016

    Article  CAS  Google Scholar 

  51. Shewmon P (1991) Diffusion in Solids, 2nd edn. Wiley-TMS, Warrendale

    Google Scholar 

  52. Srinivasan S, Desch PB, Schwarz RB (1991) Metastable phases in the Al3X (X = Ti, Zr, and Hf) intermetallic system. Scr Metall Mater 25(11):2513–2516. https://doi.org/10.1016/0956-716X(91)90059-A

    Article  CAS  Google Scholar 

  53. Vecchio KS, Williams DB (1987) Convergent beam electron diffraction study of Al3Zr in Al-Zr AND Al-Li-Zr alloys. Acta Metall 35(12):2959–2970. https://doi.org/10.1016/0001-6160(87)90295-1

    Article  CAS  Google Scholar 

  54. Doi M, Miyazaki T, Wakatsuki T (1984) The effect of elastic interaction energy on the morphology of γ′precipitates in nickel-based alloys. Mater Sci Eng 67(2):247–253. https://doi.org/10.1016/0025-5416(84)90056-9

    Article  CAS  Google Scholar 

  55. Ardell AJ (1985) Precipitation hardening. MTA 16(12):2131–2165. https://doi.org/10.1007/BF02670416

    Article  Google Scholar 

  56. Malis T, Cheng SC, Egerton RF (1988) EELS log-ratio technique for specimen-thickness measurement in the TEM. J Electron Microsc Tech 8(2):193–200. https://doi.org/10.1002/jemt.1060080206

    Article  CAS  Google Scholar 

  57. D. Ashby, “Practical statistics for medical research. Douglas G. Altman, Chapman and Hall, London, 1991. No. of pages: 611. Price: £32.00,” Statistics in Medicine, vol. 10, no. 10, pp. 1635–1636, 1991, doi: https://doi.org/https://doi.org/10.1002/sim.4780101015

  58. Balluffi RW, Allen SM, Carter WC, Kemper RA (2005) Kinetics of materials. Wiley, Hoboken

    Book  Google Scholar 

Download references

Acknowledgements

This work was supported by the Office of Naval Research (N000141612878). The authors acknowledge the Applied Chemical and Morphological Analysis Laboratory at Michigan Tech for use of the instruments and staff assistance.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yang Yang.

Ethics declarations

Conflict of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Additional information

Handling by Handling Editor: Sophie Primig.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1 TEM sample thickness determination using EELS

The relative sample thickness in the units of the local inelastic mean free path λ was determined by the ratio of zero-loss electrons intensity to the total transmitted intensity [56]:

$$t/\lambda = - \ln \left( {I_{0} /I_{t} } \right)$$
(9)

where \(I_{0}\) is the zero-loss intensity and \(I_{t}\) is the total intensity. The areas for the precipitate count and thickness measurement were chosen along the edge of the twin-jet polished holes. Since the thickness of the twin-jet polished samples increases quickly away from the edge of the hole, the sample thickness was measured at a series of points away from the edge (Fig. 

Figure 16
figure 16

a An example of sample thickness measurement for the selected area along the polished hole. b Sample thickness was measured at five places away from the edge in the spectrum image. c Spectrum of spectrum image of place #1

16).The absolute thickness was determined using the mean free path length of Al due to the small number of solute atoms. The relative and absolute thickness of the measured points were obtained according to Eq. 9 (Table 6).

The sample microstructures include precipitate size, dislocation density, and precipitate number density. These characteristics were approximately measured within areas which are less than 200 nm away from the edge.

Several, but not all, areas of dislocation density analysis had the thickness characterized by EELS thickness measurement. Areas under constant imaging conditions (near < 001 > Al) with similar gray scale contrast were also utilized for the microstructure measurement. In order to verify that the thickness approximations were consistent, the individual dislocation segment lengths in the grains were also measured. That is, because the dislocation segment lengths are limited by surface intersection, significant variations in grain thickness from sample to sample would result in significant variations in dislocation segment lengths. Such a measurement is thus a check that the assumptions of similar thicknesses (and thickness distributions) is valid. As shown in Table

Table 7 Mean dislocation length in overaged melt-spun ribbon (error bars are 95% standard error of the mean)

7, the variation in segment length between the relevant samples is varies within the standard deviation of the measurement. The significance of the mean difference between two mean values is in the range of 50%–90% (sample size equals to 10) [57].

Determining the defect density (number/volume) using the projected density (number/area) in an electropolished foil is not straight forward. Our approach has been to measure (via EELS) or estimate the thickness variation of the sample and use the following approach involving an average thickness with a linear variation in thickness from the edge of the hole to the interior of the grain. The sample thickness, \(y\), at position \(x\) is defined as:

$$y = m \cdot x + b$$
(10)

where \(m\) is the linear thickness change rate with increasing x, \(b\) is the initial thickness at \(x = 0\). The particle count depends on the sample thickness, with the defect count over a variation in thickness given by the numerator in Eq. 11 with the effective volume given by the denominator in Eq. 11, thus revealing the defect density. Note that the area term in Eq. 11 need not be specified as it algebraically cancels.

$$\frac{{{\text{Number}}\,{\text{of}}\,{\text{Defects}}}}{{{\text{Volume}}}} = \frac{{\frac{1}{L}\mathop \int \nolimits_{0}^{L} \rho \cdot \left( {m \cdot x + b} \right) \cdot {\text{area d}}x}}{{\frac{1}{L}\left[ {\mathop \int \nolimits_{0}^{L} \left( {m \cdot x + b} \right) \cdot {\text{area d}}x} \right]}} = \rho$$
(11)

To test this approximation, we examine the projected defect density as a function of distance from \(x = 0\). For all overaged melt-spun ribbon and extruded rod samples, the defect counts of precipitate and dislocation were measured in the range of 0–100 nm and 0–200 nm from hole edge into grain interior, respectively. The defect density ratios from the different counts (0–200 nm/0–100 nm) were listed in Table

Table 8 Ratios of experimentally measured precipitate area number density and dislocation area density in the range of 0–200 nm to 0–100 nm from the edge to the grain, respectively

8. According to Table 6, the measured sample thickness was 31 nm, 52 nm and 75 nm at position \(x = 0\) nm, 100 nm and 200 nm respectively.

Considering the twin-jet electropolished sample showed trapezoid-shape with linearly increasing thickness from the polished hole edge to grain (Table 6), the ratios of precipitate number and total dislocation line length are determined as:

$$\frac{{\mathop \int \nolimits_{0}^{2L} \rho \cdot \left( {m \cdot x + b} \right) \cdot {\text{area}}\,{\text{d}}x}}{{\mathop \int \nolimits_{0}^{L} \rho \cdot \left( {m \cdot x + b} \right) \cdot {\text{area}}\, {\text{d}}x}} = \frac{2 \cdot L \cdot m}{{2 \cdot b + L \cdot m}} + 2$$
(12)

where \(L = 100{\text{nm}}\) for each measurement step. According to Table 6, the measured sample thickness was 31 nm, 52 nm and 75 nm with position away from the edge at 0 nm, 100 nm and 200 nm respectively. Thus values are determined as \(m = 0.22\) and \(b = 31{\text{nm}}\). The ratio has a value of ~ 2.5, similar to the values found in the Table 8.

Appendix 2 Increasing time exponent n as a function of evolving precipitate composition

The precipitate radius \(R\) change described in traditional LSW model was defined as:

$$\frac{{{\text{d}}R}}{{{\text{d}}t}} = \frac{{2D\Omega^{2} C_{{{\text{eq}}}} \gamma }}{kTR} \cdot \left( {\frac{1}{R} - \frac{1}{R}} \right)$$
(13)

where D is the diffusivity of solute, \(\gamma\) is the interfacial energy, \(\Omega\) is the solute atomic volume, \(C_{{{\text{eq}}}}\) is the equilibrium solubility determined by the phase diagram, \(k\) is the boltzmann constant and T is the over aging temperature. The precipitate radius change depends on the size comparison between the actual precipitate size \(\left\langle R \right\rangle\) and the mean precipitate size \(R\). The time exponent of traditional LSW is 3 according to Eq. 13.

The composition of precipitates in melt-spun ribbon versus precipitate radius during annealing is plotted in Fig. 

Figure 17
figure 17

Solute concentration of precipitate in melt-spun ribbon change with precipitate radius aged from 12 to 72 h at 400 °C (error bars are 95% standard error of the mean)

17. The concentration of Sc decreased from 19.5 to 18.9% and the concentration of Zr increased from 3.1 to 3.4% with precipitate radius coarsened from 7.5 to 12.1 nm. Note that the experimentally measured compositions of precipitate are the average composition of entire precipitate. The composition change of Sc and Zr might lead more significant impact than measured values due to the inhomogeneous distribution of elements.

The confidence in these reported solutes changes was estimated using the Thiel-Sen estimator which is the median of the slopes through all points (Table

Table 9 The confidence of solutes changes for precipitate in supersaturated and dilute alloys using Thiel-Sen estimator

9). Based on the confidence of solutes changes using data from both supersaturated, the decrease of Sc composition is poorly supported with less than 30% confidence, and the increase in Zr composition shows at least 70% confidence. As for the dilute alloys [50], the decrease in Sc composition in precipitate is found to be statistically significant with 95% confidence, and the increase of Zr composition shows 85% confidence. Thus, the composition of Sc in precipitate is thought decreased and Zr is increased with annealing.

Using the lattice mismatch difference between \({\text{Al}}_{3} {\text{Zr}}\) versus \({\text{Al}}_{3} {\text{Sc}}\) (1.32% for \({\text{Al}}_{3} {\text{Sc}},\) \(0.75{\text{\% for Al}}_{3} {\text{Zr}}\)) and the concentrations changes (Table 9), the lattice mismatch for an average size precipitate decreased 1.6% after overaging from 12 to 72 h. As the interfacial energy per unit area \(\gamma\) decreases with lattice mismatch \(\varepsilon\), the interfacial energy change could be estimated in terms of radius change using the chain rule as [58]:

$$\gamma = \gamma_{0} + \left( {\frac{{{\text{d}}\varepsilon }}{{{\text{d}}C_{Zr} }} \cdot \frac{{{\text{d}}C_{Zr} }}{{{\text{d}}R}} + \frac{{{\text{d}}\varepsilon }}{{{\text{d}}C_{Sc} }} \cdot \frac{{{\text{d}}C_{Sc} }}{{{\text{d}}R}}} \right) \cdot \frac{{{\text{d}}\gamma }}{{{\text{d}}\varepsilon }} \cdot \left( {R - R_{0} } \right) = \gamma_{0} + \alpha \cdot \left( {R - R_{0} } \right)$$
(14)

where \(R_{0}\) and \(\gamma_{0}\) are the initial precipitate radius and interfacial energy and \(\alpha < 0\).Substituting Eq. 14 into Eq. 13:

$$\frac{{{\text{d}}R}}{{{\text{d}}t}} \cong \frac{{aD\Omega^{2} C_{{{\text{eq}}}} }}{kT} \cdot \left( {\gamma_{0} + \alpha \cdot \left( {R - R_{0} } \right)} \right) \cdot \left( {\frac{1}{\left\langle R \right\rangle } - \frac{1}{R}} \right) \cdot \frac{1}{R}$$
(15)

Compared to the traditional LSW theory, Eq. 15 introduces an extra term describing the change in interfacial energy with precipitate radius, \(\alpha\) (due to the change in Sc/Zr concentration with precipitate radius). Examining the specific case for \(R = 2\left\langle R \right\rangle\), which is the radius that has the maximum radial velocity in the LSW treatment:

$$\frac{{{\text{d}}R}}{{{\text{d}}t}} \cong \frac{{aD\Omega^{2} C_{{{\text{eq}}}} }}{kT} \cdot \left( {\gamma_{0} + \alpha \cdot \left( {2\left\langle R \right\rangle - R_{0} } \right)} \right) \cdot \left( {\frac{1}{{4\left\langle R \right\rangle^{2} }}} \right)$$
(16)

Integration of Eq. 16 gives a hypergeometric solution, which has a closed form Taylor series about \(\alpha\) (linear) as:

$$8\left\langle R \right\rangle^{3} \cdot \left( {\frac{1}{{3\gamma_{0} }} - \frac{\alpha }{{\gamma_{0}^{2} }}\left( {\frac{\left\langle R \right\rangle }{2} - \frac{{R_{0} }}{3}} \right)} \right) = t \cdot \frac{kT}{{aD\Omega^{2} C_{{{\text{eq}}}} }} + {\text{constant}}$$
(17)

The precipitate radius \(R\) coarsening at 400 °C described in Eq. 17 are plotted in Fig. 

Figure 18
figure 18

Precipitate radius predicted using modified LSW versus traditional LSW from t = 0 s to t = 10000 s at 400 °C

18 from 0 to 10000 s.

As the average precipitate radius calculated based on the modified LSW (Eq. 17) for \(\alpha < 0\) is found smaller than the average precipitate radius calculated based on traditional LSW at given annealing duration, the time exponent of radius for the traditional LSW was increased to larger than 3.

As the average precipitate radius calculated based on the modified LSW (Eq. 17) for \(\alpha < 0\) is found smaller than the average precipitate radius calculated based on traditional LSW at given annealing duration, the time exponent of radius for the traditional LSW was increased to larger than 3. The increased time exponent was estimated from precipitate size change in modified LSW (Table

Table 10 Four precipitate sizes were chosen from modified LSW (Fig. 18)

10).

The time exponent was found increased to 3.3 for modified LSW plot (Fig.

Figure 19
figure 19

Log–log plot of mean precipitate radii of modified LSW

19).

Appendix 3 Increasing time exponent n due to increasing precipitate volume fraction

Precipitate volume fractions were found to increase during overaging heat treatment of both the melt-spun ribbon and extruded rod. For a small increase in the total volume of the precipitates during coarsening, the traditional LSW approach can be modified into:

$$\frac{d}{{{\text{d}}t}}\mathop \sum \limits_{{{\text{part}}}} \frac{4\pi }{3}R^{3} = \delta_{1}$$
(18)

where \(\delta_{1} = 0\) in the traditional LSW derivation, but is taken as a positive constant in this modification. Rewriting Eq. 13 as:

$$\mathop \sum \limits_{{{\text{part}}}} R^{2} \frac{{{\text{d}}R}}{{{\text{d}}t}} = \delta_{2}$$
(19)

where \(\delta_{1}\) in Eq. 18 and \(\delta_{2}\) in Eq. 19 are non-zero constants. The growth rate of a particle with radius \(R\) is still considered as controlled by the mean field average concentration, \(\left\langle c \right\rangle\):

$$\frac{{{\text{d}}R}}{{{\text{d}}t}} = - D\frac{{c^{{{\text{eq}}}} - \left\langle c \right\rangle }}{R}\Omega$$
(20)

where \(D\) is the solute diffusivity, \(\Omega\) is the atomic volume, \(c^{{{\text{eq}}}}\) is the equilibrium solute concentration at the interface of a particle of radius, R, as determined by the Gibbs-Thompson effect.

Combining Eqs. 19 and 20 gives:

$$\mathop \sum \limits_{part} R\left( {c^{eq} - \left\langle c \right\rangle } \right) = - \delta$$
(21)

where \(\delta = 3\delta_{1} /\left( {4\pi D\Omega } \right)\) is a constant. Substituting the Gibbs-Thompson expression \(c^{eq} = c^{eq} \left( \infty \right) \cdot \left( {1 + \frac{2\gamma \Omega }{{kTR}}} \right)\) into Eq. 21:

$$\mathop \sum \limits_{{{\text{part}}}} R \cdot \left[ {\left\langle c \right\rangle - c^{{{\text{eq}}}} \left( \infty \right) \cdot \left( {1 + \frac{2\gamma \Omega }{{kTR}}} \right)} \right] = \delta$$
(22)

where \(c^{{{\text{eq}}}} \left( \infty \right)\) is the solute concentration from the phase diagram, \(\gamma\) is the interfacial energy, and \(k\) is the Boltzmann constant. Rearranging Eq. 22 gives:

$$\left( {\left\langle c \right\rangle - c^{{{\text{eq}}}} \left( \infty \right)} \right) \cdot \mathop \sum \limits_{{{\text{part}}}} R + \frac{{2\gamma \Omega c^{{{\text{eq}}}} \left( \infty \right)}}{kT} \cdot \mathop \sum \limits_{{{\text{part}}}} 1 = \delta$$
(23)

Equation 23 could be further rearranged gives:

$$\left( {\left\langle c \right\rangle - c^{eq} \left( \infty \right)} \right) \cdot \left\langle R \right\rangle + \frac{{2\gamma \Omega c^{eq} \left( \infty \right)}}{kT} = \frac{\delta }{{N^{tot} }}$$
(24)

where \(\left\langle R \right\rangle\) is the average particle radius, given by:

$$\left\langle R \right\rangle = \frac{{\mathop \sum \nolimits_{{{\text{part}}}} R}}{{N^{{{\text{tot}}}} }}$$
(25)

where \(N^{{{\text{tot}}}} = \mathop \sum \limits_{{{\text{part}}}} 1\) is the total number of particles. Based on Eq. 24, the average concentration \(\left\langle c \right\rangle\) is expressed as:

$$\left\langle c \right\rangle = \frac{\delta }{{N^{{{\text{tot}}}} \cdot \left\langle R \right\rangle }} + c^{{{\text{eq}}}} \left( \infty \right) \cdot \left( {1 + \frac{{2\gamma \Omega c^{{{\text{eq}}}} \left( \infty \right)}}{kT \cdot \left\langle R \right\rangle }} \right)$$
(26)

Substituting Eq. 26 into Eq. 20 gives:

$$\frac{{{\text{d}}R}}{{{\text{d}}t}} = \frac{D\Omega }{R} \cdot \left( {\frac{\delta }{{N^{{{\text{tot}}}} \left\langle R \right\rangle }} + \frac{{2\gamma \Omega c^{{{\text{eq}}}} \left( \infty \right)}}{kT\left\langle R \right\rangle } - \frac{{2\gamma \Omega c^{{{\text{eq}}}} \left( \infty \right)}}{kTR}} \right)$$
(27)

The total number of particles can be expressed in terms of equilibrium precipitate volume fraction \(f_{v}^{{{\text{eq}}}}\) as:

$$N^{{{\text{tot}}}} = \frac{3}{{4\left\langle R \right\rangle^{3} }} \cdot f_{v}^{{{\text{eq}}}}$$
(28)

Substituting Eq. 28 into Eq. 27 gives:

$$\frac{{{\text{d}}R}}{{{\text{d}}t}} = \frac{D\Omega }{R} \cdot \left( {\frac{{4\delta \left\langle R \right\rangle^{2} }}{{3f_{v}^{{{\text{eq}}}} }} + \frac{{2\gamma \Omega c^{{{\text{eq}}}} \left( \infty \right)}}{kT\left\langle R \right\rangle } - \frac{{2\gamma \Omega c^{{{\text{eq}}}} \left( \infty \right)}}{kTR}} \right)$$
(29)

Solving for \(R_{{{\text{max}}}}\) in terms of \(R\):

$$\frac{d}{dt}\left( {\frac{dR}{{dt}}} \right) = 0 = - \frac{{D\Omega \cdot \left( {4\delta R\left\langle R \right\rangle^{3} kT - 12c^{eq} \left( \infty \right)f_{v}^{eq} \left\langle R \right\rangle \gamma \Omega + 6c^{eq} \left( \infty \right)f_{v}^{eq} R\gamma \Omega } \right)}}{{3f_{v}^{eq} R^{3} \left\langle R \right\rangle kT}}$$
(30)

Yields an \(R_{{{\text{max}}}}\) of:

$$R_{{{\text{max}}}} = \frac{{6c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \left\langle R \right\rangle \gamma \Omega }}{{2\delta \left\langle R \right\rangle^{3} kT + 3c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}$$
(31)

Equation 29 is arranged by substituting \(R = R_{{{\text{max}}}}\):

$$\frac{{{\text{d}}\left\langle R \right\rangle }}{{{\text{d}}t}} = \frac{{D\Omega \cdot \left( {4\delta \left\langle R \right\rangle^{3} KT + 6c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega } \right)^{2} }}{{72c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}^{{2}} }} \left\langle R \right\rangle^{2} kT}} \cdot \left[ {\frac{{6c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}{{2\delta \left\langle R \right\rangle^{3} KT + 3c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }} - \frac{{144c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega \left\langle R \right\rangle^{3} kT}}{{\left( {4\delta \left\langle R \right\rangle^{3} KT + 6c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega } \right)^{2} }}} \right]$$
(32)

Simplify the above equation by choosing the first order of \(\delta\) gives:

$$\frac{{{\text{d}}\left\langle R \right\rangle }}{{{\text{d}}t}} = \frac{{D\Omega^{2} c^{{{\text{eq}}}} \left( \infty \right)\gamma }}{{4kT\left\langle R \right\rangle^{2} }} + \frac{{2D\Omega^{2} \left\langle R \right\rangle \gamma }}{{kTf_{v}^{{{\text{eq}}}} }}$$
(33)

Integrating Eq. 33 gives the evolution duration \(t\):

$$t = \frac{{f_{v}^{{{\text{eq}}}} \cdot \ln \left( {\left\langle R \right\rangle^{3} + \frac{{c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}{4\delta kT}} \right)}}{3D\Omega \delta } + C$$
(34)

where \(C\) is a constant that at \(t = 0\) has a value of:

$$C = - \frac{{f_{v}^{{{\text{eq}}}} \cdot \ln \left( {\left\langle {R_{0} } \right\rangle^{3} + \frac{{c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}{4\delta kT}} \right)}}{3D\Omega \delta }$$
(35)

Rearranging Eq. 34 gives:

$$t = \frac{{f_{v}^{{{\text{eq}}}} \cdot \ln \left( {\left\langle R \right\rangle^{3} + \frac{{c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}{4\delta kT}} \right)}}{3D\Omega \delta } - \frac{{f_{v} \cdot \ln \left( {\left\langle {R_{0} } \right\rangle^{3} + \frac{{c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}{4\delta kT}} \right)}}{3D\Omega \delta }$$
(36)

Equation 36 is rearranged and simplified by choosing first order of \(\delta\) to obtain:

$$t = - \frac{{4kT \cdot \left( {\left\langle R \right\rangle^{3} - \left\langle {R_{0} } \right\rangle^{3} } \right) \cdot \left( {2\delta \left\langle R \right\rangle^{3} kT + 2\delta \left\langle {R_{0} } \right\rangle^{3} kT - c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega } \right)}}{{3Dc^{{{\text{eq}}}} \left( \infty \right)^{2} f_{v}^{{{\text{eq}}}} \gamma^{2} \Omega^{2} }}$$
(37)

Equation 37 is changed into the general LSW when \(\delta = 0\):

$$t = \frac{{4kT \cdot \left( {\left\langle R \right\rangle^{3} - \left\langle {R_{0}^{3} } \right\rangle } \right) \cdot c^{{{\text{eq}}}} \left( \infty \right)f_{v}^{{{\text{eq}}}} \gamma \Omega }}{{3Dc^{{{\text{eq}}}} \left( \infty \right)^{2} f_{v}^{{{\text{eq}}}} \gamma^{2} \Omega^{2} }}$$
(38)

An experimental value for \(\delta = 3\delta_{1} /\left( {4\pi D\Omega } \right)\) is determined as \(1.2 \times 10^{14} {\text{mol}}/{\text{m}}^{2}\). The precipitate radius \(R\) coarsening at 400 °C described in Eqs. 37 and 38 are plotted in Fig.

Figure 20
figure 20

Precipitate radius predicted using modified LSW versus traditional LSW from t = 0 s to t = 10000 s at 400 °C

20 from 0 to 10000 s.

The precipitate radius calculated by modifying LSW to consider increasing precipitate volume fraction is found to be larger than the general LSW, so the time exponent of radius for the modified LSW was decreased to smaller than 3. Based on the discussions of time exponent in "Appendixs 2and 3", the values of time exponent in melt-spun ribbon and extruded rod compromised to the changed precipitate composition and increased precipitate volume fraction during annealing.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, Y., Licavoli, J.J., Hackney, S.A. et al. Coarsening behavior of precipitate Al3(Sc,Zr) in supersaturated Al-Sc-Zr alloy via melt spinning and extrusion. J Mater Sci 56, 11114–11136 (2021). https://doi.org/10.1007/s10853-021-05981-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10853-021-05981-4

Navigation