Next Article in Journal
PL Tunable GaN Nanoparticles Synthesis through Femtosecond Pulsed Laser Ablation in Different Environments
Previous Article in Journal
Flux-Free Diffusion Joining of SiCp/6063 Al Matrix Composites Using Liquid Gallium with Nano-Copper Particles in Atmosphere Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Synthesis of C-O Grain Boundary Strengthening of Al Composites

State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(3), 438; https://doi.org/10.3390/nano10030438
Submission received: 7 January 2020 / Revised: 14 February 2020 / Accepted: 25 February 2020 / Published: 29 February 2020
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)

Abstract

:
This research presents an approach for C-O grain boundary strengthening of Al composites that used an in situ method to synthesize a C-O shell on Al powder particles in a vertical tube furnace. The C-O reinforced Al matrix composites (C-O/Al composites) were fabricated by a new powder metallurgy (PM) method associated with the hot pressing technique. The data indicates that Al4C3 was distributed within the Al matrix and an O-Al solution was distributed in the grain boundaries in the strengthened structure. The formation mechanism of this structure was explained by a combination of TEM observations and molecular dynamic simulation results. The yield strength and ultimate tensile strength of the C-O/Al composites, modified by 3 wt.% polyvinyl butyral, reached 232.2 MPa and 304.82 MPa, respectively; compared to the yield strength and ultimate tensile strength of the pure aluminum processed under the same conditions, there was an increase of 124% and 99.3%, respectively. These results indicate the excellent properties of the C-O/Al-strengthened structure. In addition, the strengthening mechanism was explained by the Hall–Petch strengthening, dislocation strengthening, and solid solution strengthening mechanisms, which represented contributions of nearly 44.9%, 15.9%, and 16.6% to the total increased strength, respectively. The remaining increment was attributed to the coupled strengthening of the C and O, which contributed 20.6% to the total increase.

1. Introduction

Grain boundaries are defective in most engineering materials and control numerous mechanical, functional, and dynamic properties [1,2]. The performance of materials can be influenced by local chemical elements, such as B [3,4], C [5,6,7], N [8,9], and O [10,11,12] nonmetallic impurities, which are introduced into the grain boundaries by conventional processing methods and working environments. For example, steels can substantially increase their toughness and hardness by introducing suitable amounts of C and B elements [13]. In addition, in the presence of process, control agents and impurities (such as C, O, and H) are usually incorporated into materials produced by mechanical alloying and introducing nonmetallic compounds [14,15,16]. In addition to improving the mechanical properties, the introduction of impurities can increase GB cohesion, significantly limit grain growth, and increase the thermal stability [17,18,19]. Therefore, it is very important to design and optimize elemental segregation at the grain boundaries to improve the material properties.
The literature results show that the distribution of C and O elements at the grain boundaries of the Al phase can significantly decrease the stacking fault energy [20]. In addition, O at the grain boundaries can significantly enhance the shear strength of the Al matrix by the solution strengthening mechanism. The C element at the grain boundaries reacts with Al to generate Al4C3, which enhances the strength of the Al matrix by precipitation strengthening and dislocation strengthening [11,12,21,22]. Therefore, it is reasonable to design a C-O coupling enhancement to the Al matrix with a specific microstructure, as shown in Figure 1a. The goal is to produce a specific strengthened microstructure comprising an Al-O solution and nanoparticles of Al4C3.
According to recent reports [23,24,25,26], C/Al composites can be produced by directly introducing different carbon structures (such as carbon nanotubes, graphite powders, and activated carbon flakes) into aluminum with powder metallurgy fabrication methods; however, it is difficult to form a strong interface between aluminum and carbon upon mixing due to the low wettability between aluminum and carbon. Thus, O is often introduced to an Al matrix by oxidation in an ambient atmosphere, but it is difficult to control the amount of O upon oxidation, and the O often reacts with Al to form nanoparticles of Al2O3 [12,27]. To overcome the problems of wettability between aluminum and carbon and control the amount of O that is incorporated, an in situ reaction was designed to produce a specific strengthened microstructure comprising an Al-O solution and nanoparticles of Al4C3. The reaction included two steps as follows:
Al   +   O     AlO ( solution )
Al   +   C     Al 4 C 3
Introducing polymers onto Al particles is a very effective way to easily achieve doping of trace elements on metal surfaces. Fluoropolymers have improved the oxidation efficiency of Al particles by introducing trace nonmetallic elements that contain C, O, and F on the surface of the Al particles [28,29]. Part of the C, H, and O was effectively removed through degradation, and the species that remained were primarily C-O-F. Thus, it is reasonable to design a C-O nanoshell on Al particles by introducing polymers onto the Al particles, as shown in Figure 1b. In addition, the reaction between the Al and C-O shell should also be controlled, and reaction (1) should first be activated; otherwise, the reaction product of Al4C3 prevents the O from being incorporated into the Al matrix. Therefore, the coating polymers should be deoxygenated easily. PVB is a very common source of C and can introduce a small amount of O. Therefore, PVB is often used to coat other metals to provide a straightforward carbon source. The O element on PVB polymers is unstable and easily activates deoxygenation. The reaction process was illustrated by molecular dynamic simulation, and the simulation results suggested that reaction (1) was first activated, as shown in Figure 1c.
In our research, PVB was used to coat Al particles in a solvent, and the PVB coating on the Al particles was vacuum dried and degraded. A C-O shell was generated on the Al particles, and the modified Al particles were consolidated by hot pressing. The specific sintering mechanism was indicated by the HRTEM results and molecular dynamics (MD) simulation.

2. Materials and Methods

The raw aluminum powder (average particle size was approximately 2–3 μm, purity >99.9%, Bai Nian Ying, Zhejiang, China) was directly mixed with different polyvinyl butyral contents (1.5 wt.%, 3 wt.%, and 4.5 wt.%, Aladdin, Shanghai, China) in ethanol. The experimental procedure is shown in the schematic diagram in Figure 1. The mixture was dried in an oven at 80 °C for 8 h to ensure complete evaporation of the ethanol solvent. Then, the aluminum powder and polyvinyl butyral mixture was placed in a furnace at 480 °C for 1 h in a vacuum atmosphere to transform the organic additives into a C-O coating. The C and O content of the modified powders were performed on a CHONS analyzer (Vario EL cube, Germany). The quantity of the nanoparticles distribution of Al4C3 in the grain as well as the Al-O-Al distribution can be evaluated by the C and O content, supposing the reaction had gone completely. Cand O content of the modified powder and calculated quantity of the Al-O solution and Al4C3 in the composites is listed in Table 1. Then, the C-O coated aluminum powder was placed into a tungsten carbide (WC) dye with a diameter of 10 mm under a pressure of 30 MPa; the green compact was a cylinder with a bottom diameter of 10 mm. Finally, the compact was heated in a vacuum hot-pressing furnace at a rate of 10 °C/min and sintered at 630 °C. The microstructure of the interface was observed on a TEM machine (JEM-2100F STEM, Japan). The tensile strength of the specimens was on a universal test system (Instron-5966, Boston, MA, USA).

3. Results and Discussion

A molecular dynamics (MD) simulation of the reaction between the vinyl butyral monomer (the monomer of PVB) and Al was performed to understand our experimental observations, such as the O absorption mechanism and carbon chain aggregation. The MD simulations used a reactive force field (ReaxFF) potential for the calculations. The chemical reactions in the Al-C-H-O system can be accurately described with the ReaxFF atomistic potentials [30]. The model comprised a vinyl butyral monomer above an Al plane, as shown in Figure 2a. The system was heated to 800 K in an NVE Ensemble system. From the MD result of the reactions of the Al-C-H-O system, the O dissociation on the vinyl butyral monomer was first activated by forming two C chains, illustrated in Figure 2b,e. Then, the O was absorbed by the Al to form an Al-O solution, and the C chains dissociated and were kept away from the Al particle surface, which is shown in Figure 2c,d. The reactions can be described as follows:
C 6 H 16 O 2     O · · +   C 4 H 8 · ·
Al   +   O     AlO ( solution )
The reaction suggests that PVB is an ideal polymer to produce the final strengthening microstructure.
PVB was therefore used to modify the Al particles. The TEM images show the PVB-coated Al particles after degradation; the C-O coated Al samples and EDS results can be seen in Figure 3. The results show that the particles were spherical and monodispersed with a particle size of 2–3 μm. Due to the aggregation of the PVB during the mixing process, there were carbon-rich locations in the coating. In the case of the Al–C sample, each of the particles consisted of an Al core, carbon shell, and C-O shell coating with a thickness of approximately 10 nm, as shown in Figure 3b. The EDS spectra in Figure 3c,d show direct evidence of the existence of Al and C. In addition, the C signal can be observed in the EDS spectra in the locations that contained aggregations, further confirming that the aggregated phase was rich in carbon. The amount of O was too low to be detected in the EDS spectra.
Figure 4 shows the transmission electron microscopy topography of the sintered sample modified by 3 wt.% PVB at 630 °C. From the TEM image and EDS results of the composite in Figure 4a–c, the Al matrix was surrounded by O at the grain boundaries. To obtain a detailed understanding of the O structure, HRTEM was used to study the structure of the grain boundaries, as shown in Figure 4d. A nanoparticle with a size of 3 nm was observed in a grain boundary of the Al matrix, and the FTT electron diffraction pattern of the nanoparticle was that of Al, suggesting that the nanoparticle comprised a solid solution of Al. The C signal was detected in the TEM image in Figure 4e, the Al/Al grain boundaries were very pure, and the only interface phase that was present was a rod-like new phase. This phase was determined to be the Al4C3 phase from the electron diffraction results, and the lattice spacing was found to be 0.833 nm, which corresponds to the (003) plane of the Al4C3 crystal.
By combining the experimental results and MD simulation results, the formation mechanism of the microstructure observed herein can be explained in detail. The PVB polymer coated the surface of the Al particles and formed a core shell model. With increasing temperature, O dissociation on the vinyl butyral monomer was first activated by the formation of two C chains, and then the O was mostly absorbed into the Al crystal grain boundaries, which is shown in Figure 4d. The literature indicates that the stability of this structure is high and that O atoms did not migrate into the interior of the crystal readily [11]. Therefore, the carbon content of the external polymer residue was high. Due to the ability of transferring lone electron pairs from O to Al, the oxygen was electronegative δ-, and the aluminum was electropositive δ+. The O acted as a link between the Al particles and the external polymer, and the reduction of the O in the external polymer caused the affinity of the outer shell layer to be further reduced. Therefore, with increasing temperature, the oxygen content of the shell layer was reduced, the wettability of the outer polymer shell was gradually reduced, and the external carbon chains tended to self-aggregate, thus reacting with Al to form Al4C3 in a dispersed state.
The tensile properties of the fabricated different C-O/Al composites modified by different PVB content and reference Al materials are shown in Figure 5. From the tensile stress–strain curves of the fabricated materials in Figure 5, the yield strength (YS) and ultimate tensile strength (UTS) of the C-O/Al composites modified by 3 wt.% PVB were 228.9 MPa and 304.8 MPa, respectively. In addition, compared with the YS and UTS of the pure aluminum processed under the same conditions, 102.1 MPa and 152.9 MPa, respectively, there was an increase of 124% and 99.3%, respectively. The C-O/Al composite modified by 4.5 wt.% PVB was turned out to have a stronger YS of 242.5 MPa and a lower UTS of 260.1 MPa than the C-O/Al composite modified by 3 wt.% PVB, which was explained by brittle phases aggradation on the grain boundaries. The tensile properties of the C-O/Al composite modified by 3 wt.% PVB content were considered to be relatively high when the designed phase was in a dispersed distribution.
In the C-O/Al composite modified by 3 wt.% PVB, different mechanisms for enhancing the grain boundaries were activated to obtain the strengthening effect, including Hall–Petch strengthening ΔσHP, dislocation strengthening Δσd, and solid solution strengthening Δσss. The contributions of each mechanism to the reinforcement of the system were quantified in the present study by using a unique methodology.
The initial C-O species at the grain boundaries limited the growth of the aluminum particles. The strengthening effect is illustrated by the relation of Hall-Petch strengthening, which is shown in Equation (3) [31]:
σ HP = k y d 1 k y d 0
where d1 is the average grain size of the Al matrix after the C-O strengthening and d0 is the average grain size of the Al matrix without the presence of PVB. For the Al alloy, ky is the Hall–Petch coefficient and was equal to 0.22 MPa [32]. The value for the Hall–Petch strengthening, ΔσHP, was calculated to be 57 MPa.
For an alternative way to enhance metals, we assumed a classic distortion interaction with a solution atmosphere, and the strengthening resulting from the presence of solutes is often generally expressed as Equation (4) [11]:
σ ss =   β G ε p c q
where G is the shear modulus of the solvent, which is 26.9 GPa for Al series alloys; c is the concentration of the solute; β is an empirically determined proportionality constant related to the obstacle strength, which is often equal to 0.1; and ε is referred to as an interaction parameter. The parameter ε reduces to the misfit strain quantity ε = | ( r matrix   r solution )   /   r matrix | , where rmatrix and rsolution are the atomic radii of the matrix and solution, respectively. Parameters p and q are constants that describe the solute spacing and dislocation solute statistical mean values, respectively. The Fleischer model was used to define the constant in the model, and p = 3/2 and q = 1/2 [33]. Δσss was calculated to be 20.19 MPa.
The parameter Δσd is the strength increment due to the CTE mismatch between the matrix and the reinforcement upon cooling. Considering the dislocation enhancement discussed above, the dislocation strengthening model is given by Equation (5) [34]:
σ d =   μ b ρ
where η is a geometric constant and is equal to 1.25 for a fcc metal, μ is the shear modulus of the matrix and is equal to 2.64 × 1010 N/m2, b is the magnitude of the Burgers vector and is equal to 0.286 nm for Al, and ρ is the dislocation density [29].
Furthermore, considering the thermal mismatch at the interface of the Al4C3 reinforcements and Al matrix, the dislocation density can be calculated by Equation (6) [35]:
ρ   = 4 V α T b ( 1 V ) ( 1 t 1 + 1 t 2 + 1 t 3 )
where V is the volume fraction of the reinforcement, Δα is the difference between the thermal expansion coefficient of the matrix and reinforcement, ΔT is the difference between the sintering temperature and room temperature, b is the Burgers vector of the Al matrix, t represents the dimension of the reinforcement (t1 is the length of the reinforcement particle, t2 is the width, and t3 is the height), and Δσd was calculated to be 21.08 MPa.
In conclusion, the greatly increased tensile strength of C-O/Al composites was due to Hall–Petch strengthening ΔσHP, dislocation strengthening Δσd, and solid solution strengthening Δσss, and each contributed 44.9%, 15.9%, and 16.6%, respectively. The remaining increment may be attributed to strengthening from the C and O, which contributed 20.6% to the total strength.

4. Conclusions

The YS and UTS of the Al composites modified by 3 wt.% PVB were 228.9 MPa and 304.82 MPa, respectively; compared to the YS and UTS of pure aluminum processed under the same conditions, they experienced an increase of 124% and 99.3%, respectively. This result indicates that the C-O/Al structure provided strengthening.
The results from the analysis of the strengthened structure suggests that Al4C3 was distributed within the Al matrix with an Al-O solution present around the grain boundaries.
The formation mechanism of this structure was explained by a combination of TEM observations and molecular dynamic simulation results. The Al-O solution reaction was first activated; then, the remaining C tended to aggregate and react with Al, generating Al4C3 at the grain boundaries.
The strengthening mechanism of the microstructure was explained by the Hall–Petch, dislocation strengthening and solid solution strengthening mechanisms, which contributed 44.9%, 15.9%, and 16.6% of the total strengthening, respectively.

Author Contributions

Conceptualization, G.L. and J.H.; methodology, J.H.; software, J.H.; validation, J.Z., Y.S. and Q.S.; formal analysis, J.H.; investigation, G.L.; resources, L.Z.; data curation, J.H.; writing—original draft preparation, J.H.; writing—review and editing, J.H.; visualization, J.H.; supervision, L.Z.; project administration, L.Z.; funding acquisition, G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (Grant No. 51521001 and 51932006), the 111 Project (Grant No. B13035), and the Joint Fund (Grant No. 6141A02022255).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cantwell, P.R.; Tang, M.; Dillon, S.J.; Luo, J.; Rohrer, G.S.; Harmer, M.P. Grain boundary complexions. Acta Mater. 2014, 62, 1–48. [Google Scholar] [CrossRef]
  2. Huang, Z.; Chen, F.; Shen, Q.; Zhang, L.; Zhang, L.; Timothy, J. Uncovering the influence of common nonmetallic impurities on the stability and strength of a Σ5 (310) grain boundary in Cu. Acta Mater. 2018, 148, 110–122. [Google Scholar] [CrossRef]
  3. Anaya, J.; Rossi, S.; Alomari, M.; Kohn, E.; Tóth, L.; Pécz, B.; Karl, D.; Travis, J.; Tatyana, I.; Bradford, B. Martin Kuball a Control of the in-plane thermal conductivity of ultra-thin nanocrystalline diamond films through the grain and grain boundary properties. Acta Mater. 2016, 103, 141–152. [Google Scholar] [CrossRef] [Green Version]
  4. César, M.; Gall, D.; Guo, H. Reducing grain-boundary resistivity of copper nanowires by doping. Phys. Rev. Appl. 2016, 5, 054018. [Google Scholar] [CrossRef]
  5. HajyAkbary, F.; Sietsma, J.; Miyamoto, G.; Furuhara, T.; Santofimia, M.J. Interaction of carbon partitioning, carbide precipitation and bainite formation during the Q&P process in a low C steel. Acta Mater. 2016, 104, 72–83. [Google Scholar]
  6. Gong, P.; Palmiere, E.J.; Rainforth, W.M. Thermomechanical processing route to achieve ultrafine grains in low carbon, microalloyed steels. Acta Mater. 2016, 119, 43–54. [Google Scholar] [CrossRef] [Green Version]
  7. Zou, Y.; Xu, Y.B.; Hu, Z.P.; Gu, X.L.; Peng, F.; Tan, X.D.; Chen, S.Q.; Han, D.T.; Misra, R.D.K.; Wang, G.D. Austenite stability and its effect on the toughness of a high strength ultra-low carbon medium manganese steel plate. Mater. Sci. Eng. A 2016, 675, 153–163. [Google Scholar] [CrossRef]
  8. Nezafati, M.; Giri, A.; Hofmeister, C.; Cho, K.; Matthew, M.; Zhou, L.; Yongho, S.; Chang-Soo, K. Atomistic study on the interaction of nitrogen and Mg lattice and the nitride formation in nanocrystalline Mg alloys synthesized using cryomilling process. Acta Mater. 2016, 115, 295–307. [Google Scholar] [CrossRef]
  9. Shi, F.; Tian, P.C.; Jia, N.; Ye, Z.H.; Qi, Y.; Liu, C.M.; Li, X.W. Improving intergranular corrosion resistance in a nickel-free and manganese-bearing high-nitrogen austenitic stainless steel through grain boundary character distribution optimization. Corros. Sci. 2016, 107, 49–59. [Google Scholar] [CrossRef]
  10. Guo, J.; Duarte, M.J.; Zhang, Y.; Bachmaier, A.; Gammer, C.; Dehm, G.; Pippan, R.; Zhang, Z. Oxygen-mediated deformation and grain refinement in Cu-Fe nanocrystalline alloys. Acta Mater. 2019, 166, 281–293. [Google Scholar] [CrossRef]
  11. Tang, F.; Gianola, D.S.; Moody, M.P.; Hemker, K.J.; Cairney, J.M. Observations of grain boundary impurities in nanocrystalline Al and their influence on microstructural stability and mechanical behavior. Acta Mater. 2012, 60, 1038–1047. [Google Scholar] [CrossRef]
  12. Balog, M.; Krizik, P.; Bajana, O.; Hu, T.; Yang, H.; Julie, M.; Enrique, J. Influence of grain boundaries with dispersed nanoscale Al2O3 particles on the strength of Al for a wide range of homologous temperatures. J. Alloys Compd. 2019, 772, 472–481. [Google Scholar] [CrossRef]
  13. Ostash, O.P.; Vytvyts’kyi, V.I. Duality of the action of hydrogen on the mechanical behavior of steels and structural optimization of their hydrogen resistance. Mater. Sci. 2012, 47, 421–437. [Google Scholar] [CrossRef]
  14. Suryanarayana, C. Mechanical Alloying and Milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  15. Bhattacharya, P.; Bellon, P.; Averback, R.S.; Stephen, S.J. Nanocrystalline TiAl powders synthesized by high-energy ball milling: Effects of milling parameters on yield and contamination. J. Alloys Compd. 2004, 368, 187–196. [Google Scholar] [CrossRef]
  16. Kunčická, L.; Lowe, T.C.; Davis, C.F.; Kocich, R.; Pohludka, M. Synthesis of an Al/Al2O3 composite by severe plastic deformation. Mater. Sci. Eng. A 2015, 646, 234–241. [Google Scholar] [CrossRef]
  17. Atwater, M.A.; Scattergood, R.O.; Koch, C.C. The stabilization of nanocrystalline copper by zirconium. Mater. Sci. Eng. A 2013, 559, 250–265. [Google Scholar] [CrossRef]
  18. Detor, A.; Schuh, C. Tailoring and patterning the grain size of nanocrystalline alloys. Acta Mater. 2007, 55, 371–379. [Google Scholar] [CrossRef]
  19. Färber, B.; Cadel, E.; Menand, A.; Schmitzl, G.; Kirchheim, R. Phosphorus segregation in nanocrystalline Ni-3.6 at% P alloy investigated with the tomographic atom probe (TAP). Acta Mater. 2000, 48, 789–796. [Google Scholar]
  20. Liu, L.; Chen, J.H.; Fan, T.W.; Liu, Z.R.; Zhang, Y.; Yuan, D. The possibilities to lower the stacking fault energies of aluminum materials investigated by first-principles energy calculations. Comput. Mater. Sci. 2015, 108, 136–146. [Google Scholar] [CrossRef]
  21. Chen, B.; Jia, L.; Li, S.; Imai, H.; Takahashi, M.; Kondoh, K. In Situ Synthesized Al4C3 Nanorods with Excellent Strengthening Effect in Aluminum Matrix Composites. Adv. Eng. Mater. 2014, 16, 972–975. [Google Scholar] [CrossRef]
  22. Zhou, W.; Chen, Q.; Sui, X.; Dong, L.; Wang, Z. Enhanced thermal conductivity and dielectric properties of Al/β-SiCW/PVDF composites. Compos. Part A 2015, 71, 184–191. [Google Scholar] [CrossRef]
  23. Miracle, D.B.; Donaldson, S.L.; Henry, S.D.; Moosbrugger, C.; Anton, G.J.; Sanders, B.R.; Hrivnak, N.; Terman, C.; Kinson, J.; Muldoon, K. ASM Handbook; ASM International Materials Park: Geauga, OH, USA, 2001. [Google Scholar]
  24. Liu, Z.Y.; Xiao, B.L.; Wang, W.G.; Ma, Z.Y. Developing high-performance aluminum matrix composites with directionally aligned carbon nanotubes by combining friction stir processing and subsequent rolling. Carbon 2013, 62, 35–42. [Google Scholar] [CrossRef]
  25. Li, H.; Kang, J.; He, C.; Zhao, N.; Liang, C.; Li, B. Mechanical properties and interfacial analysis of aluminum matrix composites reinforced by carbon nanotubes with diverse structures. Mater. Sci. Eng. A-Struct. 2013, 577, 120–124. [Google Scholar] [CrossRef]
  26. Kwon, H.; Estili, M.; TakAli, K.; Miyazaki, T.; Kawasaki, A. Combination of hot extrusion and spark plasma sintering for producing carbon nanotube reinforced aluminum matrix composites. Carbon 2009, 47, 570–577. [Google Scholar] [CrossRef]
  27. Lin, Y.; Xu, B.; Feng, Y.; Lavernia, E.J. Stress-induced grain growth during high-temperature deformation of nanostructured Al containing nanoscale oxide particles. J. Alloys Compd. 2014, 596, 79–85. [Google Scholar] [CrossRef]
  28. Liu, J.; Lu, X.; Wu, C. Effect of Preparation Methods on Crystallization Behavior and Tensile Strength of Poly(vinylidene fluoride)Membranes. Membranes 2013, 3, 389–405. [Google Scholar] [CrossRef] [Green Version]
  29. Kim, K.T.; Kim, D.W.; Kim, C.K. A Facile Synthesis and Efficient Thermal Oxidation of Polytetrafluoroethylene-Coated Aluminum Powders. Mater. Lett. 2016, 167, 262–265. [Google Scholar] [CrossRef]
  30. Castro-Marcano, F.; van Duin, A.C.T. Comparison of thermal and catalytic cracking of 1-heptene from ReaxFF reactive molecular dynamics simulations. Combust. Flame 2013, 160, 766–775. [Google Scholar] [CrossRef]
  31. Ma, K.; Hu, T.; Yang, H.; Topping, T.; Yousefiani, A.; Lavernia, E.J.; Schoenung, M.J. Coupling of dislocations and precipitates: Impact on the mechanical behavior of ultrafine grained Al-Zn-Mg alloys. Acta Mater. 2016, 103, 53–164. [Google Scholar] [CrossRef]
  32. Ma, K.; Wen, H.; Hu, T.; Topping, T.D.; Isheim, D.; Seidman, D.N.; Lavernia, E.J.; Schoenung, J.M. Mechanical behavior and strengthening mechanisms in ultrafine grain precipitation-strengthened aluminum alloy. Acta Mater. 2014, 62, 141–155. [Google Scholar] [CrossRef]
  33. Fleichner, R.L. Solid solution hardening. In The Strengthening of Metals; Peckner, D., Ed.; Reinhold: New York, NY, USA, 1964; p. 93. [Google Scholar]
  34. Wu, C.; Ma, K.; Wu, J.; Luo, G.; Chen, F.; Shen, Q.; Zhang, L.; Schoenung, J.M.; Lavernia, E.J. Influence of particle size and spatial distribution of B4C reinforcement on the microstructure and mechanical behavior of precipitation strengthened Al alloy matrix composites. Mater. Sci. Eng. A 2016, 675, 421–430. [Google Scholar] [CrossRef]
  35. Arsenault, R.J.; Shi, N. Dislocation generation due to differences between the coefficients of thermal expansion. Mater. Sci. Eng. 1986, 81, 175–187. [Google Scholar] [CrossRef]
Figure 1. Schematic of the structural design and powder preparation flow chart: (a) designed structure of the C-O/Al composite, (b) designed powder structure, (c) molecular simulation of the reaction design, (d), (e), and (f) schematic diagram of the powder pretreatment.
Figure 1. Schematic of the structural design and powder preparation flow chart: (a) designed structure of the C-O/Al composite, (b) designed powder structure, (c) molecular simulation of the reaction design, (d), (e), and (f) schematic diagram of the powder pretreatment.
Nanomaterials 10 00438 g001
Figure 2. Molecular dynamics simulation of the reaction between vinyl butyral monomer and Al particles: (a) initial model of the vinyl butyral monomer and Al particles, (b) O atom dissociation process from the vinyl butyral monomer, (c) and (d) the formation of a solid solution with the Al particles, and (e) the actual process of O atom dissociation from the vinyl butyral monomer.
Figure 2. Molecular dynamics simulation of the reaction between vinyl butyral monomer and Al particles: (a) initial model of the vinyl butyral monomer and Al particles, (b) O atom dissociation process from the vinyl butyral monomer, (c) and (d) the formation of a solid solution with the Al particles, and (e) the actual process of O atom dissociation from the vinyl butyral monomer.
Nanomaterials 10 00438 g002
Figure 3. TEM images and EDS maps of the Al particles modified by 3 wt.% PVB: (a) TEM image of the Al powder modified by PVB after degradation, (b) HRTEM image of the modified Al surface, and (c) and (d) EDS mapping of the Al and C distribution.
Figure 3. TEM images and EDS maps of the Al particles modified by 3 wt.% PVB: (a) TEM image of the Al powder modified by PVB after degradation, (b) HRTEM image of the modified Al surface, and (c) and (d) EDS mapping of the Al and C distribution.
Nanomaterials 10 00438 g003
Figure 4. TEM images of the C-O/Al composites modified by 3 wt.% PVB consolidated at 630 °C: (a) TEM image of the C-O/Al composite; (b) and (c) EDS mapping of Al and O distributions, respectively; (d) HRTEM of the Al-O microstructure and its selected area electronic diffraction pattern; (e) TEM images of the Al4C3 distribution in the Al matrix; and (f) HRTEM of the Al4C3 microstructure and its selected area electronic diffraction.
Figure 4. TEM images of the C-O/Al composites modified by 3 wt.% PVB consolidated at 630 °C: (a) TEM image of the C-O/Al composite; (b) and (c) EDS mapping of Al and O distributions, respectively; (d) HRTEM of the Al-O microstructure and its selected area electronic diffraction pattern; (e) TEM images of the Al4C3 distribution in the Al matrix; and (f) HRTEM of the Al4C3 microstructure and its selected area electronic diffraction.
Nanomaterials 10 00438 g004
Figure 5. The tensile strength of the C-O/Al composite modified by different PVB contents and pure Al sintered at 630 °C.
Figure 5. The tensile strength of the C-O/Al composite modified by different PVB contents and pure Al sintered at 630 °C.
Nanomaterials 10 00438 g005
Table 1. C and O content of the modified powder and calculated quantity of Al-O solution and Al4C3 in the composites.
Table 1. C and O content of the modified powder and calculated quantity of Al-O solution and Al4C3 in the composites.
Al Modified by PVB Content (wt.%)Modified PowdersBulk Composite (Calculation Result)
C Content (wt.%)O Content (wt.%)Al4C3 Content (wt.%)Al-O Solution Content (wt.%)
1.500.5130.0191.5380.032
3.001.0360.0413.1070.069
4.501.6310.0634.8920.106

Share and Cite

MDPI and ACS Style

Hu, J.; Zhang, J.; Luo, G.; Sun, Y.; Shen, Q.; Zhang, L. Design and Synthesis of C-O Grain Boundary Strengthening of Al Composites. Nanomaterials 2020, 10, 438. https://doi.org/10.3390/nano10030438

AMA Style

Hu J, Zhang J, Luo G, Sun Y, Shen Q, Zhang L. Design and Synthesis of C-O Grain Boundary Strengthening of Al Composites. Nanomaterials. 2020; 10(3):438. https://doi.org/10.3390/nano10030438

Chicago/Turabian Style

Hu, Jianian, Jian Zhang, Guoqiang Luo, Yi Sun, Qiang Shen, and Lianmeng Zhang. 2020. "Design and Synthesis of C-O Grain Boundary Strengthening of Al Composites" Nanomaterials 10, no. 3: 438. https://doi.org/10.3390/nano10030438

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