Next Article in Journal
Effect of Cathodic Polarisation Switch-Off on the Passivity and Stability to Crevice Corrosion of AISI 304L Stainless Steel
Next Article in Special Issue
Effect of Heat Treatment Temperature on Microstructure and Properties of PM Borated Stainless Steel Prepared by Hot Isostatic Pressing
Previous Article in Journal
Improvement in the Quantification of Foreign Object Defects in Carbon Fiber Laminates Using Immersion Pulse-Echo Ultrasound
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nucleation Behavior of a Single Al-20Si Particle Rapidly Solidified in a Fast Scanning Calorimeter

1
Materials Science, Faculty of Mechanical Engineering and Marine Technology, University of Rostock, Justus-von-Liebig-Weg 2, 18059 Rostock, Germany
2
Competence Centre °CALOR, Department Life, Light & Matter, University of Rostock, Albert-Einstein-Str. 25, 18059 Rostock, Germany
3
Otto Schott Institute of Materials Research, Friedrich-Schiller-University Jena, Löbdergraben 32, 07743 Jena, Germany
4
Medical Biology and Electron Microscopy Centre, University Medicine Rostock, Strempelstr. 14, 18057 Rostock, Germany
5
Institute of Physics, University of Rostock, Albert-Einstein-Str. 23–24, 18059 Rostock, Germany
6
Butlerov Institute of Chemistry, Kazan Federal University, 18 Kremlyovskaya Street, 420008 Kazan, Russia
*
Author to whom correspondence should be addressed.
Materials 2021, 14(11), 2920; https://doi.org/10.3390/ma14112920
Submission received: 6 May 2021 / Revised: 25 May 2021 / Accepted: 27 May 2021 / Published: 28 May 2021
(This article belongs to the Special Issue Powder Processing of Light Alloys and Composites)

Abstract

:
Understanding the rapid solidification behavior characteristics, nucleation undercooling, and nucleation mechanism is important for modifying the microstructures and properties of metal alloys. In order to investigate the rapid solidification behavior in-situ, accurate measurements of nucleation undercooling and cooling rate are required in most rapid solidification processes, e.g., in additive manufacturing (AM). In this study, differential fast scanning calorimetry (DFSC) was applied to investigate the nucleation kinetics in a single micro-sized Al-20Si (mass%) particle under a controlled cooling rate of 5000 K/s. The nucleation rates of primary Si and secondary α-Al phases were calculated by a statistical analysis of 300 identical melting/solidification experiments. Applying a model based on the classical nucleation theory (CNT) together with available thermodynamic data, two different heterogeneous nucleation mechanisms of primary Si and secondary α-Al were proposed, i.e., surface heterogeneous nucleation for primary Si and interface heterogenous nucleation for secondary α-Al. The present study introduces a practical method for a detailed investigation of rapid solidification behavior of metal particles to distinguish surface and interface nucleation.

1. Introduction

In the production chains of metal components, e.g., in spray forming, rapid solidification of metal particles is a critical process. Recently, further interest has been gained in rapid solidification as an important step in additive manufacturing (AM) [1,2,3,4,5]. As important engineering materials, aluminum-based alloys, e.g., Al-Si alloys, are frequently used in AM [6,7,8,9]. In AM processes, the cooling rates are typically from ca. 100 K/s up to 106 K/s [5,10,11,12,13]. For example, in laser assisted powder bed fusion (LPBF), the cooling rate of Al-12Si (mass%) alloy was estimated via simulations, which was above 103 K/s for most parts of the melt pool [11].
Undercooling and nucleation in solidification determine many important aspects of the relevant microstructural features, e.g., the grain size, and the final material properties [14,15,16,17]. Studies regarding the nucleation undercooling of Al-Si alloy particles during rapid solidification processes has been reported [15,18,19,20,21]. However, in-situ observation of the rapid solidification behavior of metal particles is not easily available due to limitations of the measurement capability. For an accurate description of the microstructure evolution during rapid solidification, the nucleation undercooling and nucleation mechanisms are crucial, as they provide access to the optimization of the process parameters. Moreover, accurate measurements of nucleation undercooling and cooling rate, as well as accurate descriptions of rapid solidification mechanisms are required for a better understanding of rapid solidification processes of metal particles at high cooling rates.
In this study, the rapid solidification of Al-20Si (mass%) alloy was investigated by differential fast scanning calorimetry (DFSC) [22] at 5000 K/s cooling, which is in the range of the cooling rate in AM. Micro-sized Al-20Si particles with a diameter of 25 μm were used. A series of 300 nucleation events (300 heating–cooling cycles) were observed using DFSC, and the nucleation kinetics in a single particle was studied by statistical analysis. By combining microstructure and DFSC investigation, the nucleation mechanisms of the primary Si and secondary α-Al phases in a rapidly solidified single Al-20Si particle is discussed. Assuming surface and interface nucleation in a spherical cavity for Si and α-Al respectively, a modified theoretical model based on classical nucleation theory (CNT) is developed.

2. Materials and Methods

In the present work, gas-atomized Al-20Si (mass%) powder particles were used, which is purchased from Nanoval GmbH & Co. KG (Berlin, Germany). The mass fractions of alloying elements of the gas-atomized powder particles was analyzed by X-ray fluorescence spectroscopy (XRF, Bruker, Billerica, MA, USA). The composition is given in Table 1, which is close to the nominal composition.
DFSC based on the thin film sensor (UFH1, Mettler Toledo, Greifensee, Switzerland) was used to characterize the melting and solidification temperatures in-situ. As indicated in Figure 1, film-thermopiles and resistive film-heaters are located at the center of the sensor membrane (amorphous silicon-nitride membrane). One particle (Particle I) of near-spherical shape with a diameter of 25 ± 0.5 µm was chosen and deposited in the center of the measurement area (with a diameter of 90 µm) by a thin copper wire with the help of a stereomicroscope (SZX16, Olympus, Tokyo, Japan) for DFSC measurements, as shown in Figure 1. Before and after the measurements, the particle size was characterized by optical microscopy (Olympus BX41). In order to improve the heat contact between the sensor membrane and the sample, a tiny amount of silicone oil was used. Due to the high heating rate (5000 K/s) and consequent short times (less than 0.12 s) at high temperatures, it was expected that the silicone oil was stable up to high temperatures, even though the boiling temperature of silicone oil is about 250 °C at low heating rates. After the first melting/solidification cycle, a slight flattening of the particle occurred. Thus, improved thermal contact between the sensor membrane and the sample was established. Because of a static temperature offset and a dynamic temperature offset due to thermal lag, the temperature calibration and correction were performed by the eutectic temperature (577 °C) of the Al-Si alloy upon heating at different heating rates before the DFSC measurements. Details of temperature calibration and correction can be found in [23]. For the investigation of the rapid solidification of this particle (Particle I) by DFSC, the particle was heated from 37 °C to 777 °C at a rate of 5000 K/s and held for 0.01 s. This maximum heating temperature ensures a certain overheating at the high heating rate, which is ca. 93 °C higher than the liquidus temperature of Al-20Si (ca. 684 °C [24]). Afterward, the particle was quenched to 37 °C at the same cooling rate. All the DFSC measurements were performed under argon (100 kPa constant pressure).
After rapid solidification by DFSC at 5000 K/s cooling, the particle (Particle I) was transferred from the sensor membrane into an epoxy block. After careful grinding and polishing of the epoxy block, a cross-section of the particle was obtained. After that, the specimen was etched with Weck’s reagent for 4 s [25]. Details of the specimen preparation of one single micro-sized metallic particle for microstructure characterization can be found in [26]. Scanning electron microscopy (SEM) (Zeiss Merlin VP Compact, Oberkochen, Germany) was used to characterize the microstructure (cross-section) of the particle.
To fully characterize the solidification behavior of the particle undergoing rapid solidification at 5000 K/s, another particle (Particle II) with the same size (25 ± 0.5 µm in diameter) as the particle for microstructure characterization (Particle I) was heated and cooled with the same time–temperature profile for 300 times. Note that the temperature calibration and correction were performed before these repeated heating–cooling measurements. Then, the nucleation undercooling, ΔT, for each scan was calculated by using the following equation:
Δ T = T l T s _ o n s e t c o r r e c t
where Tl is the liquidus temperature and T s _ o n s e t c o r r e c t is the solidification onset temperature (nucleation temperature). The evaluation of T s _ o n s e t c o r r e c t can be found in [23].

3. Results and Discussion

The microstructure of the cross-section of a single Al-20Si particle (Particle I) rapidly solidified at 5000 K/s cooling by DFSC was characterized by SEM, as shown in Figure 2. The grey and bright areas in Figure 2 are Al and Si, respectively. The microstructure is comprised of the primary Si, secondary α-Al dendrites, and (α-Al + Si) eutectics. It seems very likely that the primary Si with polygonal morphology solidified from the surface of the particle (see Figure 2). Unlike the microstructure of gas-atomized particles, only two Si grains were observed because of the lower cooling rate than that of gas-atomized particles.
Figure 3 shows the corresponding DFSC curve at the heating–cooling rate of 5000 K/s for Particle I shown in Figure 2. The primary Si phase starts to solidify first during rapid solidification of an Al-20Si alloy, followed by the formation of the α-Al phase, and then the (α-Al + Si) eutectic. Two solidification onsets upon cooling (as indicated by the arrows in Figure 3) correspond to the starts of the nucleation of the primary Si and the secondary α-Al. The corresponding nucleation temperatures of the Si and α-Al phases for Particle I cooled at 5000 K/s are 627 °C and 521 °C, respectively. Note that the nucleation temperature of the α-Al + Si eutectics cannot be evaluated precisely because of the overlapping solidification peaks (at ca. 500 °C) between α-Al dendrites and α-Al + Si eutectics, as shown in Figure 3. For the nucleation of the primary Si, the nucleation undercooling can be evaluated by Equation (1) with the liquidus temperature of the Al-20Si alloy (684 °C), which is 57 K. After the nucleation of primary Si, primary Si grows and rejects Al into the melt with decreasing temperature. At 521 °C (lower than the eutectic temperature 577 °C [24]), Al nucleates on the interface of primary Si and starts to grow in the melt. The melt is depleted of Si because of the solidification of primary Si. Assuming that the concentration of Si in the melt in front of the solidifying Si follows the metastable extension of the Al liquidus line in the Al-Si phase diagram [24] during the rapid solidification of primary Si, the concentration of Si is 9% (mass%) at 521 °C. The liquidus temperature at this concentration is 597 °C [24], and thus, the nucleation undercooling of α-Al is 76 K.
For a sound statistical basis, another particle (Particle II) with the same size (25 ± 0.5 μm in diameter) as Particle I was heated and cooled at 5000 K/s for 300 times. Figure 4 shows the nucleation temperatures and undercoolings of the Si and α-Al phases for each of the 300 nucleations in chronological order. Both nucleations occur in bands of approximately 30 K width because of the stochastic nature of nucleation. The nucleation temperatures of the primary Si and α-Al for Particle I are also in these bands.
If nucleation events are independent and occur randomly in time as described in CNT, the events fulfill the requirements for a Poisson process. These requirements can be fulfilled for the nucleation process of single metallic particles [16,27,28]. Assuming one nucleus forms in one nucleation event at the solidification onset temperature, the nucleation rate per unit volume and unit surface area, JV and JS, can be expressed as [29,30]:
J S = λ β S
and:
J V = λ β V
where V and S are the volume and surface area of the melt, β is the cooling rate, and λ is the time-dependent rate constant of an inhomogeneous Poisson distribution, which can be determined from the histogram distributions of Si and α-Al (see the insets of Figure 5) using the following equation [27]:
λ = a b ( s + a / 2 )
where b is the width of the temperature interval (=1 K in this study). In one of the temperature intervals, a nucleation events occur and s nucleation events remain molten to enter the next temperature interval. As mentioned above, the surface nuclei of the Si phase form at the surface of the particle (S = πd2, 1963 ± 80 μm2, where d is the particle diameter), and the nucleation rate of the primary Si is calculated by Equation (2), as shown in Figure 5a.
After solidification of the primary Si, the volume of the melt is reduced and the residual Si in the melt is less than the initial melt concentration. Hence, the volume of the melt and the amount of residual Si in the melt should be estimated. In the absence of the solidification enthalpy of Si in the deep undercooled region, the melting enthalpy of Si is used as its solidification enthalpy at ca. 627 °C. Thus, the mass of the solidified Si, mSi, is calculated by the following expression:
m S i = Δ H s _ D F S C Δ H m _ S i
where ΔHs_DFSC is the solidification enthalpy measured by DFSC (2.4 ± 0.2 μJ for the DFSC curve shown in Figure 3) and ΔHm_Si is the melting enthalpy of Si (1726 J/g) [31]. It should be pointed out that ΔHs_DFSC was evaluated by integrating the Si solidification shoulder (peak) until α-Al starts to solidify, as shown in Figure 3. It is assumed that the solidification of Si is limited by the nucleation events of α-Al. Thus, the mass of the solidified Si, mSi_s, can be calculated and equals 1.4 ± 0.1 ng, corresponding to a volume of 600 ± 40 μm3. Then, assuming that the nuclei for α-Al form on the primary Si (interface nucleation on the primary Si), whose volume is 7581 ± 480 μm3, the nucleation rate for α-Al is calculated by Equation (3), as shown in Figure 5b. As expected, both nucleation rates increase with increasing undercooling. It is supposed that the size of the primary Si and α-Al decrease with increasing undercooling.
To understand the tendency between the nucleation rate and undercooling in terms of the spherical cap model of CNT, the expressions for the surface and interface heterogeneous nucleation rates of a single particle during rapid solidification are derived in [32,33] as:
J S = n S D l a 0 σ l s k ( T m _ o n s e t Δ T ) exp ( Δ G * f S k ( T m _ o n s e t Δ T ) )
and:
J V = n V D l a 0 σ l s k ( T m _ o n s e t Δ T ) exp ( Δ G * f V k ( T m _ o n s e t Δ T ) )
with:
Δ G * = 16 π σ l s 3 3 Δ G V 2
where nS and nV are the densities of surface and interface heterogeneous nucleation sites (the numbers of potential surface and interface heterogeneous nucleation sites per unit surface area and unit volume), respectively, Dl is the liquid diffusivity, a0 is the atomic spacing, σls is the solid-liquid interfacial energy (surface tension), k is the Boltzmann constant, fS and fV are the shape factors for surface and interface heterogeneous nucleation, and ∆GV is the free energy difference between the melt and the solidified phase. The shape factors describe the potency of the heterogeneous nuclei.
During the DFSC measurements, an inherent Al2O3 layer always exists, while the primary Si is solidified from the surface of the particle, as shown in Figure 2. Therefore, it is assumed that the surface nuclei for Si form in a spherical cavity at the surface of the particle (surface nucleation on the Al2O3 oxide layer). For simplicity of the model, it is also assumed that the interface nuclei for α-Al form in a spherical cavity (at a nanosized scale) inside the particle (interface nucleation on the primary Si). Thus, the shape factors, fS and fV is given by [34]:
f S , V = 1 4 ρ S , V 3 ( 1 + 2 ρ S , V + 1 + ρ S , V 2 + 2 ρ S , V cos θ S , V ) ( 1 + ρ S , V 1 + ρ S , V 2 + 2 ρ S , V cos θ S , V ) 2
with:
ρ S , V = r * R S , V
where subscript S,V designates the surface and interface heterogeneous nucleation, respectively, θS,V is the contact angle between the heterogeneous nucleus and the undercooled melt, RS,V is the radius of the spherical cavity, and r* is the critical radius of a nucleus. Based on CNT, r* is given by:
r * = 2 σ l s Δ G V
Due to the lack of specific heat capacity data for each phase, especially in the undercooled region, ∆GV is estimated by a parallel tangent construction [35]. Based on a thermodynamic assessment of the Al-Si system, the Gibbs energies of solid Al, solid Si, and liquid phases are represented by polynomial expansions [24]:
G i = G 0 A l i ( 1 x ) + G 0 S i i x + R T [ x ln x + ( 1 x ) ln ( 1 x ) ] + x ( 1 x ) [ A i + B i ( 1 2 x ) + C i ( 1 6 x + 6 x 2 ) ]
where i designates the phase, x is the atomic fraction of Si (0.19 and 0.09 for the nucleation of Si and Al, respectively), R is the gas constant, T is absolute temperature, G 0 A l i and G 0 S i i are the lattice stability terms of the pure Al and Si, and Ai, Bi, and Ci are the interaction parameters that are linearly dependent on temperature. Table 2 lists the parameters for Equation (12).
The development of a complete description of the density of nucleation sites within an undercooled melt has been a continuing challenge. As mentioned above, the primary Si nuclei form on the surface of the particle. Therefore, it is assumed that the surface heterogeneous nucleation of Si could occur on the whole Al2O3 layer on the surface of the Al-20Si particle. Thus, the density of surface heterogeneous nucleation sites for primary Si can be expressed by [23]:
n S = N A h V m
where NA is the Avogadro constant, h is the thickness of the Al2O3 layer (=1 nm) [36], and Vm is the molar volume of Al2O3 (=25.7 × 10−6 m3/mol) [36]. Thus, nS equals 2.343 × 1019 m−2. For the interface nucleation of α-Al, the density of interface heterogeneous nucleation sites in the melt is given by [15]:
n V = n V 0 exp ( c Δ T )
where n V 0 is the base level of nucleation sites (=1013 m−3) and c is a positive constant (=0.059) [15].
Based on this model (Equations (6)–(14)), the relationship between the nucleation rate and nucleation undercooling for the rapid solidification of single particles with a certain heterogeneity can be estimated. Employing the obtained parameters listed in Table 3, the data were fitted for Si and α-Al phases. Here θS, θV, RS, and RV are fitting parameters. Levenberg-Marquardt nonlinear least-squares curve fitting was utilized with boundary conditions of 20° < θS,V ≤ 180° and 0.1 nm < RS,V < 10 nm. For the spherical cap model, the contact angle is larger than 20° [37]. The fitted curves are indicated by dashed lines in Figure 5. The best fits to the data yield RS = 1.29 nm and θS = 74.6° for the surface nucleation of Si, and RV = 2.61 nm and θV = 53.4° for the interface nucleation of α-Al. As shown in Figure 5, the modeling of the nucleation rates based on CNT with the inhomogeneous Poisson analysis of the thermal cycling behavior gives an accurate description of the 300 nucleation measurements for the single micro-sized Al-20Si particle that is rapidly solidified at 5000 K/s. Moreover, the nucleation mechanisms for the primary Si and secondary α-Al are well described as surface and interface heterogeneous nucleation in a spherical cavity.

4. Conclusions

In-situ DFSC was successfully used to investigate the rapid solidification process of a single micro-sized Al-20Si (mass%) particle at a cooling rate of 5000 K/s. By DFSC treatments at the same cooling rate, nucleation undercooling of one single micro-sized particle was correlated with the microstructure analysis of another similarly particle (with the same size). The results show that the microstructure of the rapidly solidified particle consists of primary Si, secondary α-Al, and (α-Al + Si) eutectics.
The nucleation rates of the primary Si and α-Al dendrites were obtained by a statistical analysis of 300 solidification events (300 heating–cooling cycles of one single particle by DFSC). According to a modified classical heterogeneous nucleation theory, two different nucleation mechanisms of the primary Si and secondary α-Al were proposed, i.e., surface heterogeneous nucleation for Si and interface heterogeneous nucleation for α-Al.
By controlled rapid cooling and known nucleation undercooling via in-situ DFSC with microstructure characterization, it is possible to investigate the rapidly solidified structures and the nucleation mechanisms of single metallic particles. Therefore, this is a practical approach for a detailed investigation of the rapid solidification processes of metal particles and to discriminate surface and interface nucleation.

Author Contributions

Conceptualization, B.Y. and O.K.; methodology, Q.P. and B.Y.; software, B.Y.; validation, B.M., M.R., C.S. and O.K.; formal analysis, Q.P., B.Y. and D.L.; investigation, Q.P. and B.Y.; resources, B.M., A.S. and O.K.; data curation, Q.P. and B.Y.; writing—original draft preparation, B.Y.; writing—review and editing, Q.P., B.M., D.L., M.R. and C.S.; visualization, Q.P., B.Y. and A.S.; supervision, O.K.; project administration, B.Y., M.R. and O.K.; funding acquisition, B.Y., M.R. and O.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was founded by the German Research Foundation (DFG), grant numbers YA417/4-1 and KE616/29-1, and the German Research Foundation (DFG) priority program SPP2122, grant numbers RE1261/23-1 and KE616/27-1.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Acknowledgments

B.Y. acknowledges support from the German Research Foundation (Project YA417/4-1). M.R. acknowledges support from the German Research Foundation (Project RE1261/23-1, SPP 2122). C.S. acknowledges financial support from the Ministry of Education and Science of the Russian Federation grant 14.Y26.31.0019. O.K. acknowledges support from the German Research Foundation (Project KE616/29-1 and KE616/27-1, SPP2122). The authors also thank Bilal Gökce from the University of Duisburg-Essen for XRF analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Konrad, C.; Zhang, Y.; Xiao, B. Analysis of melting and resolidification in a two-component metal powder bed subjected to temporal Gaussian heat flux. Int. J. Heat Mass Transf. 2005, 48, 3932–3944. [Google Scholar] [CrossRef]
  2. Konrad, C.; Zhang, Y.; Shi, Y. Melting and resolidification of a subcooled metal powder particle subjected to nanosecond laser heating. Int. J. Heat Mass Transf. 2007, 50, 2236–2245. [Google Scholar] [CrossRef]
  3. Gusarov, A.V.; Yadroitsev, I.; Bertrand, P.; Smurov, I. Heat transfer modelling and stability analysis of selective laser melting. Appl. Surf. Sci. 2007, 254, 975–979. [Google Scholar] [CrossRef]
  4. Gu, D.D.; Meiners, W.; Wissenbach, K.; Poprawe, R. Laser additive manufacturing of metallic components: Materials, processes and mechanisms. Int. Mater. Rev. 2012, 57, 133–164. [Google Scholar] [CrossRef]
  5. Li, Y.; Gu, D. Parametric analysis of thermal behavior during selective laser melting additive manufacturing of aluminum alloy powder. Mater. Des. 2014, 63, 856–867. [Google Scholar] [CrossRef]
  6. Delahaye, J.; Tchuindjang, J.T.; Lecomte-Beckers, J.; Rigo, O.; Habraken, A.M.; Mertens, A. Influence of Si precipitates on fracture mechanisms of AlSi10Mg parts processed by Selective Laser Melting. Acta Mater. 2019, 175, 160–170. [Google Scholar] [CrossRef]
  7. Kleiner, S.; Zürcher, J.; Bauer, O.; Margraf, P. Heat treatment response of selectively laser melted AlSi10Mg. HTM 2020, 75, 327–341. [Google Scholar] [CrossRef]
  8. Prashanth, K.G.; Scudino, S.; Klauss, H.J.; Surreddi, K.B.; Löber, L.; Wang, Z.; Chaubey, A.K.; Kühn, U.; Eckert, J. Microstructure and mechanical properties of Al–12Si produced by selective laser melting: Effect of heat treatment. Mater. Sci. Eng. A 2014, 590, 153–160. [Google Scholar] [CrossRef]
  9. Aboulkhair, N.T.; Simonelli, M.; Parry, L.; Ashcroft, I.; Tuck, C.; Hague, R. 3D printing of Aluminium alloys: Additive Manufacturing of Aluminium alloys using selective laser melting. Prog. Mater. Sci. 2019, 106, 100578. [Google Scholar] [CrossRef]
  10. Doubenskaia, M.A.; Zhirnov, I.V.; Teleshevskiy, V.I.; Bertrand, P.; Smurov, I.Y. Determination of true temperature in selective laser melting of metal powder using infrared camera. Mater. Sci. Forum. 2015, 834, 93–102. [Google Scholar] [CrossRef]
  11. Li, X.P.; Wang, X.J.; Saunders, M.; Suvorova, A.; Zhang, L.C.; Liu, Y.J.; Fang, M.H.; Huang, Z.H.; Sercombe, T.B. A selective laser melting and solution heat treatment refined Al–12Si alloy with a controllable ultrafine eutectic microstructure and 25% tensile ductility. Acta Mater. 2015, 74–82. [Google Scholar] [CrossRef]
  12. Farshidianfar, M.H.; Khajepour, A.; Gerlich, A.P. Effect of real-time cooling rate on microstructure in Laser Additive Manufacturing. J. Mater. Process Technol. 2016, 231, 468–478. [Google Scholar] [CrossRef]
  13. Hooper, P.A. Melt pool temperature and cooling rates in laser powder bed fusion. Addit. Manuf. 2018, 22, 548–559. [Google Scholar] [CrossRef]
  14. Trivedi, R. The role of heterogeneous nucleation on microstructure evolution in peritectic systems. Scr. Mater. 2005, 53, 47–52. [Google Scholar] [CrossRef]
  15. Gremaud, M.; Allen, D.R.; Rappaz, M.; Perepezko, J.H. The development of nucleation controlled microstructures during laser treatment of Al-Si alloys. Acta Mater. 1996, 44, 2669–2681. [Google Scholar] [CrossRef]
  16. Wilde, G.; Sebright, J.L.; Perepezko, J.H. Bulk liquid undercooling and nucleation in gold. Acta Mater. 2006, 54, 4759–4769. [Google Scholar] [CrossRef]
  17. Gandin, C.-A.; Mosbah, S.; Volkmann, T.; Herlach, D.M. Experimental and numerical modeling of equiaxed solidification in metallic alloys. Acta Mater. 2008, 56, 3023–3035. [Google Scholar] [CrossRef]
  18. Trivedi, R.; Jin, F.; Anderson, I.E. Dynamical evolution of microstructure in finely atomized droplets of Al-Si alloys. Acta Mater. 2003, 51, 289–300. [Google Scholar] [CrossRef]
  19. Das, S.K.; Perepezko, J.H.; Wu, R.I.; Wilde, G. Undercooling and glass formation in Al-based alloys. Mater. Sci. Eng. A 2001, 304–306, 159–165. [Google Scholar] [CrossRef]
  20. Perepezko, J.H.; Sebright, J.L.; Höckel, P.G.; Wilde, G. Undercooling and solidification of atomized liquid droplets. Mater. Sci. Eng. A 2002, 326, 144–153. [Google Scholar] [CrossRef]
  21. Perepezko, J.H.; LeBeau, S.E.; Mueller, B.A.; Hildeman, G.J. Rapid solidification of highly undercooled aluminum powders. In Rapidly Solidified Powder Aluminum Alloys; Fine, M.E., Starke, E.A., Eds.; ASTM International: West Conshohocken, PA, USA, 1986; pp. 118–136. ISBN 978-0-8031-4961-8. [Google Scholar]
  22. Zhuravlev, E.; Schick, C. Fast scanning power compensated differential scanning nano-calorimeter: 1. The device. Thermochim Acta 2010, 50, 1–13. [Google Scholar] [CrossRef]
  23. Yang, B.; Peng, Q.; Milkereit, B.; Springer, A.; Liu, D.; Rettenmayr, M.; Schick, C.; Keßler, O. Nucleation behaviour and microstructure of single Al-Si12 powder particles rapidly solidified in a fast scanning calorimeter. J. Mater. Sci. 2021, 56, 12881–12897. [Google Scholar] [CrossRef]
  24. Murray, J.L.; McAlister, A.J. The Al-Si (Aluminum-Silicon) system. Bull. Alloy. Phase Diagr. 1984, 5, 74–84. [Google Scholar] [CrossRef]
  25. Weck, E.; Leistner, E. Metallographic Instructions for Colour Etching by Immersion, Part III: Non-Ferrous Metals, Cemented Carbides and Ferrous Metals, Nickel-Base and Cobalt-Base Alloys; Deutscher Verlag für Schweißtechnik: Düsseldorf, Germany, 1986. [Google Scholar]
  26. Milkereit, B.; Meißner, Y.; Ladewig, C.; Osten, J.; Yang, B.; Springer, A.; Keßler, O. Metallographische Präparation einzelner Pulver-Partikel. Prakt. Metallogr. 2021, 58, 129–139. [Google Scholar] [CrossRef]
  27. Uttormark, M.J.; Zanter, J.W.; Perepezko, J.H. Repeated nucleation in an undercooled aluminum droplet. J. Cryst. Growth 1997, 177, 258–264. [Google Scholar] [CrossRef]
  28. Yang, B.; Gao, Y.; Zou, C.; Zhai, Q.; Zhuravlev, E.; Schick, C. Repeated nucleation in an undercooled tin droplet by fast scanning calorimetry. Mater. Lett. 2009, 63, 2476–2478. [Google Scholar] [CrossRef]
  29. Bokeloh, J.; Rozas, R.E.; Horbach, J.; Wilde, G. Nucleation barriers for the liquid-to-crystal transition in Ni: Experiment and simulation. Phys. Rev. Lett. 2011, 107, 145701. [Google Scholar] [CrossRef] [PubMed]
  30. Simon, C.; Gao, J.; Mao, Y.; Wilde, G. Fast scanning calorimetric study of nucleation rates and nucleation transitions of Au-Sn alloys. Scr. Mater. 2017, 139, 13–16. [Google Scholar] [CrossRef]
  31. NIST. Chemistry WebBook: NIST Standard Reference Database Number 69; Linstrom, P.J., Mallard, W.G., Eds.; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2021. [Google Scholar]
  32. Turnbull, D. Formation of crystal nuclei in liquid metals. J. Appl. Phys. 1950, 21, 1022–1028. [Google Scholar] [CrossRef]
  33. Gutzow, I.; Schmelzer, J.W.P. The Vitreous State: Thermodynamics, Structure, Rheology, and Crystallization; Springer: Berlin, Germany, 1995. [Google Scholar]
  34. Iwamatsu, M. Line-tension effects on heterogeneous nucleation on a spherical substrate and in a spherical cavity. Langmuir 2015, 31, 3861–3868. [Google Scholar] [CrossRef] [Green Version]
  35. Thompson, C.V.; Spaepen, F. Homogeneous crystal nucleation in binary metallic melts. Acta Metall. 1983, 31, 2021–2027. [Google Scholar] [CrossRef]
  36. Korb, L.J.; Olson, D.L. ASM Handbook: Corrosion; ASM International: Materials Park, OH, USA, 1992. [Google Scholar]
  37. Cantor, B. Heterogeneous nucleation and adsorption. Philos. Trans. R. Soc. A 2003, 361, 409–417. [Google Scholar] [CrossRef]
  38. Kalay, Y.E.; Chumbley, L.S.; Anderson, I.E.; Napolitano, R.E. Characterization of hypereutectic Al-Si powders solidified under far-from equilibrium conditions. Metall. Mater. Trans. A 2007, 38, 1452–1457. [Google Scholar] [CrossRef] [Green Version]
  39. Gündüz, M.; Hunt, J.D. The measurement of solid-liquid surface energies in the Al-Cu, Al-Si and Pb-Sn systems. Acta Metall. 1985, 33, 1651–1672. [Google Scholar] [CrossRef]
  40. Faraji, M. The Effect of Solidification Variables on the Microstructure of Hypereutectic Al-Si Alloys. Ph.D. Thesis, The University of Sheffield, Sheffield, UK, 2007. [Google Scholar]
Figure 1. Single Al-20Si powder particle (Particle I) on a UFH1 DFSC sensor.
Figure 1. Single Al-20Si powder particle (Particle I) on a UFH1 DFSC sensor.
Materials 14 02920 g001
Figure 2. SEM images (secondary electron images) of the cross-section of an Al-20Si particle (Particle I) solidified at 5000 K/s, etched with Weck’s reagent.
Figure 2. SEM images (secondary electron images) of the cross-section of an Al-20Si particle (Particle I) solidified at 5000 K/s, etched with Weck’s reagent.
Materials 14 02920 g002
Figure 3. DFSC heat flow curve of a single Al-20Si particle (Particle I) at the cooling rate of 5000 K/s. The melting and solidification onset temperatures of different phases are indicated by arrows.
Figure 3. DFSC heat flow curve of a single Al-20Si particle (Particle I) at the cooling rate of 5000 K/s. The melting and solidification onset temperatures of different phases are indicated by arrows.
Materials 14 02920 g003
Figure 4. Nucleation temperatures and undercoolings of primary Si and α-Al dendrites as measured from 300 cycles.
Figure 4. Nucleation temperatures and undercoolings of primary Si and α-Al dendrites as measured from 300 cycles.
Materials 14 02920 g004
Figure 5. Nucleation rates as a function of undercooling for the (a) primary Si and (b) α-Al dendrites. The dashed curves indicate the best fit to the data for the surface and interface heterogeneous nucleation models. Distributions of nucleation events versus undercooling are shown in the insets.
Figure 5. Nucleation rates as a function of undercooling for the (a) primary Si and (b) α-Al dendrites. The dashed curves indicate the best fit to the data for the surface and interface heterogeneous nucleation models. Distributions of nucleation events versus undercooling are shown in the insets.
Materials 14 02920 g005
Table 1. Mass fractions of alloying elements in the gas-atomized Al-20Si powder particles.
Table 1. Mass fractions of alloying elements in the gas-atomized Al-20Si powder particles.
AlloySiFeMgAgTiCuAl
Al-20Si20%0.27%0.07%0.03%0.02%0.02%Bal.
Table 2. Thermodynamic model parameters, J/mol, T in K [15,24].
Table 2. Thermodynamic model parameters, J/mol, T in K [15,24].
G 0 A l f c c = 10,792 + 11.56 T A l = 10,695.4 1.823 T
G 0 S i f c c = 12.12 T B l = 4275.5 + 3.044 T
G 0 S i d i a = 50,600 + 30.00 T C l = 670.7 0.460 T
G 0 A l d i a = 30.00 T A f c c = 200 7.594 T
G 0 A l l = 0 A d i a = 89,138 31.445 T
G 0 S i l = 0
Table 3. Parameters used for nucleation rate analysis.
Table 3. Parameters used for nucleation rate analysis.
ParametersValue
Liquidus temperature (Si), TlSi684 °C [24]
Liquidus temperature (Al), TlAl597 °C [24]
Liquid diffusivity, Dl5 × 10−9 m2/s [38]
Interfacial energy (Al in Al-Si system), σslAl0.169 J/m2 [39]
Interfacial energy (Si in Al-Si system), σslSi0.352 J/m2 [39]
Atomic spacing (Al), a0Al2.6 × 10−10 m (pure Al) [27]
Atomic spacing (Si), a0Si3.0 × 10−10 m (Si atomic diameter) [40]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, Q.; Yang, B.; Milkereit, B.; Liu, D.; Springer, A.; Rettenmayr, M.; Schick, C.; Keßler, O. Nucleation Behavior of a Single Al-20Si Particle Rapidly Solidified in a Fast Scanning Calorimeter. Materials 2021, 14, 2920. https://doi.org/10.3390/ma14112920

AMA Style

Peng Q, Yang B, Milkereit B, Liu D, Springer A, Rettenmayr M, Schick C, Keßler O. Nucleation Behavior of a Single Al-20Si Particle Rapidly Solidified in a Fast Scanning Calorimeter. Materials. 2021; 14(11):2920. https://doi.org/10.3390/ma14112920

Chicago/Turabian Style

Peng, Qin, Bin Yang, Benjamin Milkereit, Dongmei Liu, Armin Springer, Markus Rettenmayr, Christoph Schick, and Olaf Keßler. 2021. "Nucleation Behavior of a Single Al-20Si Particle Rapidly Solidified in a Fast Scanning Calorimeter" Materials 14, no. 11: 2920. https://doi.org/10.3390/ma14112920

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop