Next Article in Journal
Research on Pre-Compensation and Shape-Control Optimization of Hemming Structures with Dissimilar Materials Based on Forming Process Chain
Previous Article in Journal
Hydrogeochemical Facies and Health Hazards of Fluoride and Nitrate in Groundwater of a Lithium Ore Deposit Basin
Previous Article in Special Issue
Effect of Electromagnetic Field Assistance on the Wear and Corrosion Resistance of Nickel-Based Coating by Laser Cladding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding the High-Temperature Deformation Behaviors in Additively Manufactured Al6061+TiC Composites via In Situ Neutron Diffraction

1
Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
2
Department of Mechanical Engineering, University of Wisconsin-Madison, Madison, WI 53706, USA
*
Author to whom correspondence should be addressed.
Metals 2024, 14(9), 1064; https://doi.org/10.3390/met14091064
Submission received: 30 July 2024 / Revised: 6 September 2024 / Accepted: 11 September 2024 / Published: 17 September 2024

Abstract

:
Aluminum matrix composites (AMCs) are designed to enhance the performance of conventional aluminum alloys for engineering applications at both room and elevated temperatures. However, the dynamic phase-specific deformation behavior and load-sharing mechanisms of AMCs at elevated temperatures have not been extensively studied and remain unclear. Here, in situ neutron diffraction experiments are employed to reveal the phase-specific structure evolution of additively manufactured Al6061+TiC composites under compressive loading at 250 °C. It is found that the addition of a small amount of nano-size TiC significantly alters the deformation behavior and increases the strength at 250 °C in comparison to the as-printed Al6061. Unlike the two-stage behavior observed in Al6061, the Al6061+TiC composites exhibit three stages during compression triggered by changes in the interphase stress states. Further analysis of Bragg peak intensity and broadening reveals that the presence of TiC alters the dislocation activity during deformation at 250 °C by influencing dislocation slip planes and promoting dislocation accumulation. These findings provide direct experimental observations of the phase-specific dynamic process in AMCs under deformation at an elevated temperature. The revealed mechanisms provide insights for the future design and optimization of high-performance AMCs.

1. Introduction

Aluminum alloys have long been essential for the automotive, aerospace, marine, and other industries due to their high strength-to-weight ratio, low cost, and excellent corrosion resistance [1,2,3]. However, common aluminum alloys often exhibit degraded mechanical performance at elevated temperatures due to their low melting points, thermal softening, microstructural instability, and low creep resistance [4,5,6]. Particle-reinforced aluminum matrix composites (AMCs) [7,8,9,10], which combine aluminum alloys as the matrix and secondary ceramic particles as the reinforcement phase, are designed to augment the performance capabilities of conventional Al alloys, particularly in environments where traditional aluminum alloys fall short, such as those at elevated temperatures. Unlike common high-strength Al alloys (e.g., Al2000, Al6000, and Al7000 series), which rely on solid-state metastable precipitation to achieve high strength [11,12,13,14], AMCs use thermodynamically stable ceramic particles as reinforcement, which enable robust strengthening at both room and elevated temperatures [15,16,17,18].
Numerous studies have demonstrated the strengthening effects in AMCs at both room and elevated temperatures by adding ceramic particles, such as TiC [19,20], SiC [21,22], Al2O3 [23,24], and others [25,26]. Adding particles activates several strengthening mechanisms to enhance the mechanical strengths of AMCs, including Orowan strengthening, the Hall–Petch effect, load sharing, etc. [27,28,29]. Load sharing between phases can exhibit complex multistage behavior throughout the deformation process, as observed in other dual or multiphase metal alloys [30,31]. In addition to load sharing between phases, the deformation behavior of the Al matrix differs significantly at elevated temperatures in terms of dislocation annihilation, the activation of additional dislocation slip systems, and precipitation-induced chemistry changes [32,33]. Adding particles can impede the dislocation motion of the Al matrix, potentially leading to more complex and distinctly different deformation behaviors in AMCs at elevated temperatures.
However, most previous research has focused on the collective response of bulk AMCs [10,34,35,36]. The dynamic phase-specific response at the grain level, especially during elevated-temperature loading, has not been well studied. This gap hinders a deeper understanding of deformation and load-sharing mechanisms as well as the design of AMCs with enhanced mechanical performance at elevated temperatures. Investigating the load-sharing and deformation behavior of AMCs at elevated temperatures requires a highly complex in situ setup, including mechanical loading, sample heating, and neutron or synchrotron diffraction. However, most previous studies on metal matrix composites using neutron or synchrotron X-rays have focused on room-temperature deformation [37]. The load-sharing and deformation behaviors of AMCs at elevated temperatures have not been extensively studied due to the limited availability of high-flux and high-resolution neutron and synchrotron sources [38,39,40].
In fusion-based additive manufacturing (AM) processes [41,42,43,44], the unique rapid heating and cooling cycles with localized high melt pool temperatures and rapid solidification rates [45,46,47,48,49] may significantly influence interphase bonding, dislocation density, and solidification texture. These factors potentially affect the load sharing, deformation behaviors, and mechanical properties of AMCs. In this work, additively manufactured Al6061+TiC composites were fabricated by the laser powder bed fusion (LPBF) process and used as a model material for in situ neutron diffraction studies to understand the phase-specific deformation behaviors at an elevated temperature. Neutrons’ capability of interacting with lightweight elements such as carbon as well as their deep-penetrating capability to measure the bulk sample on average give rise to a high-quality signal of the Al matrix, a small amount of TiC, and minor precipitate phases from the alloy, clearly characterizing the dynamic and kinetic evolution of phase stability, lattice strain, and phase-specific stress under combined sample environments, such as thermomechanical testing at elevated temperatures.

2. Materials and Methods

2.1. Materials

Al6061, Al6061+2 vol.%TiC, and Al6061+5 vol.%TiC were fabricated by the laser powder bed fusion (LPBF) AM process. The Al6061 powders used in the LPBF process were purchased from Valimet (Stockton, CA, USA). The detailed chemical composition of the Al6061 powders is shown in Table 1. The Al6061+2 vol.%TiC and Al6061+5 vol.%TiC powders were prepared by the planetary ball milling of Al6061 powders and TiC nanoparticles (83 nm, SSNano, Houston, TX, USA). Alumina grinding jars with a capacity of 100 mL and alumina grinding balls with a diameter of 6 mm were used for this study. The jars were loaded with 20 g of Al6061+TiC powder, maintaining a ball-to-powder weight ratio of 5:1. The disc rotation speed was set to 200 RPM, with a 10 min pause after every 20 min milling cycle. The total milling time was 8 h.
The LPBF process employed a pulsed laser beam with a Gaussian beam profile. The detailed laser processing parameters are listed in Table 2. Printing was conducted using the stripe scan pattern [50], with the scanning vector rotated by 67° following each layer. The printed sample had dimensions of 15 mm (length), 7 mm (width), and 20 mm (height).

2.2. In Situ Neutron Diffraction

In situ neutron diffraction experiments were conducted at the VULCAN diffractometer, Spallation Neutron Source, Oak Ridge National Laboratory [51], to investigate the load-sharing and deformation behavior of Al6061 and Al6061+TiC AMCs at the elevated temperature of 250 °C. Figure 1 shows the experiment setup, and more details on instrument layout and optics can be found in reference [51]. The cuboid specimen, measuring 12 mm (length) × 6 mm (width) × 6 mm (height), was initially heated to 250 °C using induction coils and maintained at this temperature for over ten minutes to allow the temperature to stabilize. Subsequently, uniaxial compression tests were performed along the sample’s length using the MTS load frame until approximately 10% strain was reached or strain hardening ceased. At the same time, time-of-flight neutron diffraction was continuously conducted on a gauge volume of 5 mm × 5 mm × 6 mm. The neutron gauge volume was defined by neutron incident slits (with a 5 mm horizontal opening and a 6 mm vertical opening) and 5 mm receiving collimators (Figure 1b). The incident neutron beam was directed at a 45° angle relative to the loading direction. The diffracted neutron beams at −90° (Bank 1) and +90° (Bank 2) relative to the incident beam were captured by two neutron detectors to obtain diffraction data from the (hkl) crystal planes along the loading direction (LD) and the normal direction (ND). The sample building direction in the LPBF was aligned with the ND in the in situ neutron diffraction experiment. A K-type thermocouple was spot-welded to the bottom center of the sample (Figure 1a), which was used to monitor and control sample temperature via the PID controller. To investigate the effect of test temperature on deformation behavior, three additional Al6061 and Al6061+TiC samples were tested at room temperature for comparison with the results obtained at 250 °C.

2.3. In Situ Neutron Data Analysis

The in situ neutron data collected during the continuous loading and unloading were chopped into 5 min time bins using VDRIVE software (https://web.ornl.gov/~kean/VDRIVE/VDRIVE.html, accessed on 1 July 2024) [52]. Single-peak fitting (SPF) for the hkl peaks of both the Al and TiC phases was performed for each data bin using VDRIVE software to obtain hkl-specific diffraction peak properties. Rietveld refinement of the full diffraction patterns was conducted for each data bin using GSAS with EXPGUI (https://subversion.xray.aps.anl.gov/trac/EXPGUI, accessed on 1 July 2024) [53,54] to obtain the phase-averaged lattice information, the secondary Mg2Si phase fraction, and the site occupancies of TiC.
The lattice strain evolution was calculated using two methods: (1) a method based on the hkl-specific d-spacing determined from the SPF, and (2) a method based on the average lattice parameter of each phase determined from the Rietveld refinement:
ε = d d 0 d 0                      
where d is the d-spacing of hkl lattice planes or the lattice parameter of the Al or TiC phases during loading and d 0 is the reference d-spacing or lattice parameter before loading. The effects of thermal residual stress on d0 determination were not considered, as previous studies have shown that no thermal residual stress exists between the Al and reinforcement phases at temperatures of 200 °C or higher [55].
On the basis of the lattice strain and Hooke’s law, the phase-specific stress was calculated by [56]:
σ p , i , 11 = E i 1 + ν i 1 2 ν i 1 ν i ε i , 11 + ν i   ( ε i , 22 + ε i , 33 )
where i represents TiC or Al, σ p , i , 11 is the phase-specific stress along LD, E i is the diffraction elastic constant (i.e., the elastic constant determined by the slope of the macroscopic stress versus the elastic lattice strain curve in the elastic stage during in situ neutron diffraction), and ν i is Poisson’s ratio. Specifically, E A l   = 68 GPa, E T i C   = 203 GPa, and ν A l = 0.31, ν T i C   = 0.22. ε i , 11 is the lattice strain along LD, ε i , 22 and ε i , 33 are the lattice strains along the ND and the transverse direction (TD). The ND and TD are two axes perpendicular to LD in the 3D Cartesian coordinate system. The lattice strain along the TD is assumed to be the same as that along the ND.
The hydrostatic stress was calculated by [57]:
σ h y , i = σ p , i , 11 + σ p , i , 22 + σ p , i , 33 3                    
where σ p , i , 11 , σ p , i , 22 , and σ p , i , 33 are the phase-specific stress along the LD, ND, and TD, respectively. The deviatoric stress was calculated by [57]:
σ d , i , 11 = σ p , i , 11 σ h y , i                              
where σ p , i , 11 is the phase-specific stress and σ h y , i is the hydrostatic stress.

3. Results and Discussion

3.1. Engineering Stress–Strain

Figure 2 presents the compressive engineering stress–strain curves for the Al6061 and Al6061+TiC composites at 250 °C. The addition of TiC significantly enhances the yield strength, which increases from 66 MPa for Al6061 to 96 MPa for Al6061+2%TiC and further to 104 MPa for Al6061+5%TiC. Moreover, the hardening behavior during the plastic deformation is also altered with TiC presence. After unloading and reloading, the Al6061 sample exhibits significant softening, whereas the Al6061+TiC samples demonstrate continuous hardening at larger strains. The macroscopic curves indicate different microscopic mechanisms with and without TiC, which is explained later via analysis of in situ neutron diffraction. Although unloading and reloading were conducted in the experiment, the elastic behavior in these steps is not discussed in the following analysis of lattice strains and phase stresses, and the corresponding data points are not plotted in the figures for simplicity.

3.2. Phase Stability and Lattice Strains

With various solid solutes in the Al matrix as well as dispersive ceramics, the high temperature possibly destabilizes the solid solution and drives the atom diffusion between phases during mechanical testing. This chemistry change may lead to a different d 0 value in Equation (1), and it can thus influence the accurate determination of elastic lattice strain and phase-specific stress via the in situ diffraction approach [56]. It is thus essential to understand the phase stability first. Neutron diffraction probes the matrix Al phase, TiC ceramic, and minor precipitates, such as Mg2Si [13] (Figure 3a), which benefit from its unique capabilities of deep penetration in a bulk-scale volume and the high sensitivity of lightweight elements. The evolution of the diffraction patterns during compression at 250 °C is visualized in Figure 3b, which reflects the phase-specific deformation behaviors and the phase stability.
Rietveld refinement was conducted to quantify the weight fraction evolution of the TiC and Mg2Si in Al6061+5%TiC during deformation at 250 °C. Both the peak height in the contour plot of neutron diffraction (Figure 4a) and the weight fraction via refinement (Figure 4b) reflect a stable TiC weight fraction of ~8%. This value agrees with the nominal TiC composition (5% by volume or 8.7% by weight). The slight decrease in the TiC weight fraction after printing may be due to some TiC particles attaching to the ball milling media and thus being slightly lost during the ball milling process. By contrast, the weight fraction of Mg2Si initially increases from 0.4% to 0.9% and then stabilizes (Figure 4b). This suggests that Mg2Si does not fully precipitate during the LPBF process and continues to precipitate during the compressive test at 250 °C.
Rietveld refinement was further used to study the possible changes in the chemical composition of the TiC phase. During the LPBF process, the atom’s interdiffusion between phases may be activated. This could result in some Al or other solutes dissolving into TiC by partially replacing Ti in the as-printed AMCs. Here, the model formula (Ti1−xAlx)C is employed in the Rietveld refinement by giving freedom of composition variation at the cation site. On the basis of this, the “occupancy” of the Al at the Ti-site during the compressive test of Al6061+5%TiC at 250 °C was calculated to indicate the chemical stability of the TiC phase. The results reveal a small “occupancy” of Al in TiC, and the value did not significantly change at 250 °C (Figure 4c). Thus, the occupancy refinement again supports the conclusion of the chemically stable TiC ceramic phase at this elevated temperature in addition to the support from the unchanged weight fraction.
Therefore, the Al lattice strain during compression at 250 °C may result from both the mechanical deformation and the time-dependent chemical changes, while the TiC lattice strain solely reflects the mechanical response. In the analysis of phase stress in the following, the deviatoric components of stress are used to deconvolute the effects of chemistry variation in the phase [56]. Although precipitates like Mg2Si are probed in neutron diffraction, the trivial fraction (<1%) is not considered to significantly impact the composite behavior over the TiC strengthening. The following analysis focuses on the Al matrix and the TiC phase.
To uncover the phase-specific mechanical responses, lattice strains for different lattice planes in the Al and TiC phases during the compressive loading of Al6061 and Al6061+TiC composites at 250 °C were determined through single-peak fitting of neutron diffraction patterns. In Al6061, the lattice strains across all Al lattice planes show an approximately linear increase throughout the loading process, encompassing both the elastic and plastic stages (Figure 5a). The divergence in lattice strains between different lattice planes is minimal, indicating low elastic and plastic anisotropy at 250 °C.
In Al6061+TiC composites, both the Al and TiC phases initially exhibit an approximately linear increase in lattice strain with applied stress (Figure 5b,c). The rate of increase is higher for the Al phase than for TiC, reflecting a higher elastic modulus of TiC than of Al. At the applied stresses of approximately −108 MPa for Al6061+2%TiC and −123 MPa for Al6061+5%TiC, the lattice strain for Al abruptly decreases, while the lattice strain for TiC continues to increase. This sudden change can be attributed to the yielding of the Al matrix, which transfers more load from the softer Al matrix to the harder TiC particles. The minimal divergence in lattice strain between different lattice planes for both Al and TiC phases indicates low elastic and plastic anisotropy in both phases. Given this low anisotropy, the average phase-specific lattice strains were determined from the changes in the average lattice parameter of each phase, which were calculated via Rietveld refinement of the full diffraction patterns. The comparisons between the average lattice strains and the {311} lattice strains via SPF are shown in Figure 5d–f and are consistent with each other. These average lattice strains were used for further phase stress analysis.

3.3. Phase-Specific Stress

The Al and TiC phase-specific stresses were directly calculated on the basis of the average lattice strain obtained from Rietveld refinement and Hooke’s law (Equation (2)) without consideration of the chemical changes. These “phase stresses” are plotted in Figure 6. In Al6061, the “stress” of the Al phase increases nearly linearly along the LD while remaining at a low minimal value along the ND (Figure 6a). This result aligns with the expected triaxial stress calculation during the uniaxial loading of a single-phase component.
In the Al6061+TiC composites (Figure 6b,c), the “stress” of both the Al and TiC phases initially shows a continuous increase with applied stress along the LD. Following the yielding of the Al matrix, the rate of “stress” increase in the Al phase decelerates in Al6061+2%TiC and even decreases in Al6061+5%TiC. In the TiC phase, the “stress” exhibits a higher rate of increase after the Al matrix yields. In addition to the overall trend, fluctuations were observed in “phase stress”. As discussed above, this may result from both the mechanical response and the chemical changes at the elevated temperature simultaneously. To deconvolute these two types of responses, the deviatoric and hydrostatic stress components of the phase-specific stress were calculated. The deviatoric component suppressed the effects of d 0 variation due to chemistry, temperature, and so forth, and this reflects the pure mechanical response in this complex environment; the hydrostatic component may reflect the comprehensive effects of all these factors [56].
Deviatoric stress is responsible for shearing and distortion in a material, which is directly related to yielding and plastic deformation. Adding TiC decreases the deviatoric stress of the Al phase. At an applied stress of −100 MPa, the deviatoric stress values of the Al phase in Al6061, Al6061+2%TiC, and Al6061+5%TiC are −73 MPa, −66 MPa, and −63 MPa, respectively. The reduction in deviatoric stress due to the presence of TiC particles delays the yielding of the Al6061+TiC composite, which explains the strength enhancement in the macro strain–stress curve in Figure 2.
Compared with the “phase stress”, the deviatoric stress shows a very smooth change in response upon applied stress (Figure 6d–f). This enables the detection of slight but important details. The Al6061+5%TiC sample that has a better signal of the TiC phase clearly exhibits three stages during the deformation, as distinguished by the slopes of the Al and TiC phases (Figure 6f). At the beginning of loading, both phases deformed elastically, showing the same deviatoric stress. After about −75 MPa of applied stress, another distinct stage is observed before the obvious plastic deformation in which the slopes differ between the phases. The TiC phase shows higher deviatoric stress than the Al phase. This may be attributed to the initiation of localized plastic deformation of the Al phase near TiC particles, leading to the early yielding of the material. The curve divergence increases upon loading, and it is further amplified in the third stage, in which the plastic deformation occurs over the Al phase globally. It is clearly observed that the existence of TiC shares the deviatoric stress and significantly reduces the deviatoric stress in the Al matrix in the plastic region under high applied stress. This can greatly delay the failure occurrence in the Al phase. In the Al6061+2%TiC sample, the slopes of the deviatoric stresses exhibit trends consistent with those in the Al6061+5%TiC sample, although the data of TiC have larger fitting error bars (Figure 6e).
The “hydrostatic stress” of the Al phase initially starts at zero and becomes increasingly negative in all samples (Figure 6g–i), indicating a continuous increase in compressive hydrostatic stress during loading. Compared to the deviatoric stress, the “hydrostatic stress” of Al may result from both the average hydrostatic stress state throughout the Al–TiC phase boundary to the depth of the Al grains and the change in chemical composition due to precipitate formation (Figure 4a,b). Differently, as the TiC phase is chemically stable, the “hydrostatic stress” reflects the stress states of the TiC particles. Irregular fluctuations are observed in Al6061+5%TiC under stress values of −75 to −120 MPa (Figure 6i), showing a complex hydrostatic stress change. This agrees with the possible interphase stress changes due to the initial plastic deformation in Al near the phase boundary, aligning with the second stage of deviatoric stress evolution.

3.4. Effects of TiC Particles on Dislocation Activity at 250 °C

Adding TiC not only affects load sharing but also alters the deformation behavior of the Al matrix in terms of dislocation activity and texture development. To assess texture evolution, changes in the peak intensities of different Al lattice planes in Al6061 and Al6061+TiC samples during loading at room temperature and 250 °C were quantified and compared (Figure 7).
At room temperature, the peak intensity of Al {220} increases, while those of Al {111} and Al {200} decrease (Figure 7a–c). This pattern is typical for aluminum alloys loaded at room temperature, in which only the {111} slip plane is activated. At elevated temperatures, the slip on the {101} plane is activated as well, in addition to that on the {111} plane [33]. The new slip system is thought to result in an increase in {200} peak intensity along the compression direction, which is the opposite of the result of the slip system at room temperature. The competition of these two systems mitigates the changes in Al {200} intensity during the deformation of as-printed Al6061 at 250 °C. Specifically, at a strain of 5%, the peak intensity of Al {200} decreases by 1.6% at 250 °C, compared with a 13.6% decrease at room temperature in Al6061 (Figure 7d). Interestingly, this competition can be altered by adding TiC particles. Figure 7e,f show a further increase in the peak intensity of Al {200} in the composites, with a 1.5% increase at a strain of 5% for Al6061+5%TiC. This suggests that TiC may inhibit the dislocation slip on the {111} plane more than that on the {101} plane at 250 °C.
Adding TiC particles also affects the amplitude of the peak intensities. At both room temperature and 250 °C, the amplitude of the peak intensity change decreases as the volume fraction of TiC increases from 2% to 5%. This effect likely occurs because more TiC particles hinder dislocations around them to strengthen the matrix instead of allowing dislocations to slip through the whole grain and develop deformation textures.
While they hinder the dislocation movement, TiC particles contribute to the accumulated dislocations in the Al matrix during the deformation, overcoming the annihilation of dislocations at 250 °C. As the higher dislocation density broadens the diffraction peaks, the peak width of different Al lattice planes during loading at 250 °C is quantified and plotted in Figure 8. In Al6061, the peak width shows an overall decreasing trend after a slight initial increase (Figure 8a), indicating that dislocation generally exhibits annihilation [58]. By contrast, the peak width in the Al6061+TiC samples shows a continuous increase after an initial rapid rise (Figure 8b,c), suggesting overall dislocation multiplication with the presence of TiC. This difference arises because TiC particles can pin dislocations in the Al matrix, promoting dislocation accumulation around the TiC particles. This accumulation counteracts dislocation annihilation, resulting in an overall increase in dislocation density in Al6061+TiC. This also explains the continuous hardening observed in the Al6061+TiC samples, in contrast to the softening observed in Al6061 (Figure 2).

4. Conclusions

In this work, Al6061+TiC composites were additively fabricated by laser powder bed fusion (LPBF) and used as a model material for in situ neutron diffraction studies to elucidate phase-specific deformation behaviors at an elevated temperature. In situ neutron diffraction was demonstrated to capture the structural evolution of the Al matrix, dispersive TiC phase, and minor precipitate from the alloy during the compression test at 250 °C. It was found that the addition of a small amount of nano-size TiC significantly increases the strength and alters the deformation behavior at 250 °C. The deviatoric stresses calculation excluded the influence of chemistry changes on the stress analysis, and it revealed the three load-sharing stages in the phases. Further analysis of Bragg peak intensity and broadening reveals that the presence of TiC altered the dislocation activity during deformation at 250 °C by influencing dislocation slip planes and promoting dislocation accumulation. The key findings are summarized below:
(1)
Adding TiC significantly enhances yield strength, which increases from 66 MPa for Al6061 to 96 MPa for Al6061+2%TiC and further to 104 MPa for Al6061+5%TiC at 250 °C.
(2)
Unlike the two-stage elastic–plastic behavior observed in Al6061, Al6061+TiC composites exhibit three stages during compression: (1) the elastic deformation of both the Al and TiC phases; (2) the early yielding of the Al matrix triggered by local plastic deformation near TiC particles, and (3) the global plastic deformation of the Al matrix.
(3)
Minor chemical changes due to Mg2Si precipitation and the dissolution of Al into TiC were identified. These changes lead to fluctuations in the measured “phase stresses” during loading at 250 °C. The addition of TiC reduces the deviatoric stress carried by the Al matrix, delaying the yielding of the metal matrix and thereby enhancing the composite’s strength.
(4)
Peak intensity quantification revealed that in addition to the {111} slip plane typically activated at room temperature, the {101} slip system is also activated at 250 °C. The addition of TiC inhibits dislocation slip on the {111} plane more than on the {101} plane at 250 °C.
(5)
The peak width of different Al lattice planes shows an overall increasing trend in Al6061+TiC, contrasting with the overall decreasing trend in pure Al6061 at 250 °C. This difference arises because TiC particles pin dislocations in the Al matrix, promoting dislocation accumulation around TiC particles and thereby increasing dislocation density in Al6061+TiC.
These findings provide direct experimental observations of the phase-specific dynamic process in AMCs under deformation at an elevated temperature. The revealed mechanisms provide insights for the future design and optimization of high-performance AMCs.

Author Contributions

Conceptualization, Y.C. and M.Q.; Methodology, Y.C., M.Q., D.Y., L.C. and K.A.; Formal analysis, M.Q.; Investigation, Y.C., M.Q., D.Y., L.C. and K.A.; Resources, Y.C., L.C. and K.A.; Funding acquisition: Y.C., L.C. and K.A.; Supervision, Y.C., L.C. and K.A.; Project Administration, Y.C.; Writing—original draft preparation, M.Q.; Writing—review and editing, Y.C., M.Q., L.C. and K.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by U.S. Department of Energy under Contract No. DE-AC05-00OR22725, and National Science Foundation grant number CMMI-2002840.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

This research used resources at the Spallation Neutron Source, a DOE Office of Science User Facility operated by Oak Ridge National Laboratory. The beam time was allocated to VULCAN with proposal number IPTS-31933. M. Qu would like to acknowledge support from the Materials Engineering Initiative of the Neutron Scattering Division, Oak Ridge National Laboratory.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hattori, C.; Almeida, G.; Gonçalves, R.; Santos, R.; Souza, R.; da Silva, W.; Cunali, J.; Couto, A. Microstructure and fatigue properties of extruded aluminum alloys 7046 and 7108 for automotive applications. J. Mater. Res. Technol. 2021, 14, 2970–2981. [Google Scholar] [CrossRef]
  2. Omiyale, B.O.; Olugbade, T.O.; Abioye, T.E.; Farayibi, P.K. Wire arc additive manufacturing of aluminium alloys for aerospace and automotive applications: A review. Mater. Sci. Technol. 2022, 38, 391–408. [Google Scholar] [CrossRef]
  3. Li, S.; Yue, X.; Li, Q.; Peng, H.; Dong, B.; Liu, T.; Yang, H.; Fan, J.; Shu, S.; Qiu, F.; et al. Development and applications of aluminum alloys for aerospace industry. J. Mater. Res. Technol. 2023, 27, 944–983. [Google Scholar] [CrossRef]
  4. Mills, K.C. Recommended Values of Thermophysical Properties for Selected Commercial Alloys; Woodhead Publishing: Sawston, UK, 2002. [Google Scholar] [CrossRef]
  5. Lee, W.-S.; Tang, Z.-C. Relationship between mechanical properties and microstructural response of 6061-T6 aluminum alloy impacted at elevated temperatures. Mater. Des. 2014, 58, 116–124. [Google Scholar] [CrossRef]
  6. Rokni, M.; Zarei-Hanzaki, A.; Abedi, H. Microstructure evolution and mechanical properties of back extruded 7075 aluminum alloy at elevated temperatures. Mater. Sci. Eng. A 2011, 532, 593–600. [Google Scholar] [CrossRef]
  7. Wang, Y.; Zhang, J. A Review of the friction and wear behavior of particle-reinforced aluminum matrix composites. Lubricants 2023, 11, 317. [Google Scholar] [CrossRef]
  8. Karabulut, Y.; Ünal, R. Additive manufacturing of ceramic particle-reinforced aluminum-based metal matrix composites: A review. J. Mater. Sci. 2022, 57, 19212–19242. [Google Scholar] [CrossRef]
  9. Chen, B.; Xi, X.; Tan, C.; Song, X. Recent progress in laser additive manufacturing of aluminum matrix composites. Curr. Opin. Chem. Eng. 2020, 28, 28–35. [Google Scholar] [CrossRef]
  10. Tang, S.; Ummethala, R.; Suryanarayana, C.; Eckert, J.; Prashanth, K.G.; Wang, Z. Additive manufacturing of aluminum-based metal matrix composites—A review. Adv. Eng. Mater. 2021, 23, 2100053. [Google Scholar] [CrossRef]
  11. Li, J.-F.; Peng, Z.-W.; Li, C.-X.; Jia, Z.-Q.; Chen, W.-J.; Zheng, Z.-Q. Mechanical properties, corrosion behaviors and microstructures of 7075 aluminium alloy with various aging treatments. Trans. Nonferrous Met. Soc. China 2008, 18, 755–762. [Google Scholar] [CrossRef]
  12. Bardel, D.; Perez, M.; Nelias, D.; Deschamps, A.; Hutchinson, C.; Maisonnette, D.; Chaise, T.; Garnier, J.; Bourlier, F. Coupled precipitation and yield strength modelling for non-isothermal treatments of a 6061 aluminium alloy. Acta Mater. 2014, 62, 129–140. [Google Scholar] [CrossRef]
  13. Tsaknopoulos, K.; Walde, C.; Champagne, V.; Cote, D. Gas-atomized Al 6061 powder: Phase identification and evolution during thermal treatment. JOM 2019, 71, 435–443. [Google Scholar] [CrossRef]
  14. Cheng, S.; Zhao, Y.; Zhu, Y.; Ma, E. Optimizing the strength and ductility of fine structured 2024 Al alloy by nano-precipitation. Acta Mater. 2007, 55, 5822–5832. [Google Scholar] [CrossRef]
  15. Qu, M.; Guo, Q.; Escano, L.I.; Nabaa, A.; Hojjatzadeh, S.M.H.; Young, Z.A.; Chen, L. Controlling process instability for defect lean metal additive manufacturing. Nat. Commun. 2022, 13, 1079. [Google Scholar] [CrossRef]
  16. Lin, T.-C.; Cao, C.; Sokoluk, M.; Jiang, L.; Wang, X.; Schoenung, J.M.; Lavernia, E.J.; Li, X. Aluminum with dispersed nanoparticles by laser additive manufacturing. Nat. Commun. 2019, 10, 4124. [Google Scholar] [CrossRef]
  17. Cao, C.; Yao, G.; Jiang, L.; Sokoluk, M.; Wang, X.; Ciston, J.; Javadi, A.; Guan, Z.; De Rosa, I.; Xie, W.; et al. Bulk ultrafine grained/nanocrystalline metals via slow cooling. Sci. Adv. 2019, 5, eaaw2398. [Google Scholar] [CrossRef] [PubMed]
  18. Qu, M.; Yuan, J.; Nabaa, A.; Huang, J.; Chuang, C.A.; Chen, L.; Qu, M.; Yuan, J.; Nabaa, A.; Huang, J.; et al. Melting and solidification dynamics during laser melting of reaction-based metal matrix composites uncovered by in-situ synchrotron X-ray diffraction. Acta Mater. 2024, 271, 119875. [Google Scholar] [CrossRef]
  19. Kaftelen, H.; Ünlü, N.; Göller, G.; Öveçoğlu, M.L.; Henein, H. Comparative processing-structure–property studies of Al–Cu matrix composites reinforced with TiC particulates. Compos. Part A Appl. Sci. Manuf. 2011, 42, 812–824. [Google Scholar] [CrossRef]
  20. Sahoo, P.; Koczak, M. Microstructure-property relationships of in situ reacted TiC/AlCu metal matrix composites. Mater. Sci. Eng. A 1991, 131, 69–76. [Google Scholar] [CrossRef]
  21. Soltani, M.; Atrian, A. High temperature tensile behavior and microstructure of Al-SiC nanocomposite fabricated by mechanical milling and hot extrusion technique. Mater. Res. Express 2018, 5, 025026. [Google Scholar] [CrossRef]
  22. Song, M. Effects of volume fraction of SiC particles on mechanical properties of SiC/Al composites. Trans. Nonferrous Met. Soc. China 2009, 19, 1400–1404. [Google Scholar] [CrossRef]
  23. Ganguly, P.; Poole, W.; Lloyd, D. Deformation and fracture characteristics of AA6061-Al2O3 particle reinforced metal matrix composites at elevated temperatures. Scr. Mater. 2001, 44, 1099–1105. [Google Scholar] [CrossRef]
  24. Park, B.G.; Crosky, A.G.; Hellier, A.K. Material characterisation and mechanical properties of Al2O3-Al metal matrix composites. J. Mater. Sci. 2001, 36, 2417–2426. [Google Scholar] [CrossRef]
  25. Li, X.; Ji, G.; Chen, Z.; Addad, A.; Wu, Y.; Wang, H.; Vleugels, J.; Van Humbeeck, J.; Kruth, J. Selective laser melting of nano-TiB2 decorated AlSi10Mg alloy with high fracture strength and ductility. Acta Mater. 2017, 129, 183–193. [Google Scholar] [CrossRef]
  26. Gao, C.; Wang, Z.; Xiao, Z.; You, D.; Wong, K.; Akbarzadeh, A. Selective laser melting of TiN nanoparticle-reinforced AlSi10Mg composite: Microstructural, interfacial, and mechanical properties. J. Mech. Work. Technol. 2020, 281, 116618. [Google Scholar] [CrossRef]
  27. Nardone, V.C.; Prewo, K.M. On the strength of discontinuous silicon carbide reinforced aluminum composites. Scr. Met. 1986, 20, 43–48. [Google Scholar] [CrossRef]
  28. Ramakrishnan, N. An analytical study on strengthening of particulate reinforced metal matrix composites. Acta Mater. 1996, 44, 69–77. [Google Scholar] [CrossRef]
  29. Clyne, T.W.; Withers, P.J. An Introduction to Metal Matrix Composites; Cambridge University Press: Cambridge, UK, 1993. [Google Scholar] [CrossRef]
  30. Wanner, A.; Dunand, D.C. Synchrotron X-ray study of bulk lattice strains in externally loaded Cu-Mo composites. Met. Mater. Trans. A 2000, 31, 2949–2962. [Google Scholar] [CrossRef]
  31. Bacon, D.; Edwards, L.; Moffatt, J.; Fitzpatrick, M. Synchrotron X-ray diffraction measurements of internal stresses during loading of steel-based metal matrix composites reinforced with TiB2 particles. Acta Mater. 2011, 59, 3373–3383. [Google Scholar] [CrossRef]
  32. Zhang, J.-S. High Temperature Deformation and Fracture of Materials; Elsevier: Amsterdam, The Netherlands, 2010. [Google Scholar] [CrossRef]
  33. Stout, M.G.; Chen, S.R.; Kocks, U.F.; Schwartz, A.J.; MacEwen, S.R.; Beaudoin, A.J. Mechanisms Responsible for Texture Development in a 5182 Aluminum Alloy Deformed at Elevated Temperature; No. LA-UR-98-2646; CONF-981054-; Los Alamos National Lab. (LANL): Los Alamos, NM, USA, 1998. [Google Scholar]
  34. Casati, R.; Vedani, M. Metal matrix composites reinforced by nano-particles—A review. Metals 2014, 4, 65–83. [Google Scholar] [CrossRef]
  35. Dadbakhsh, S.; Mertens, R.; Hao, L.; Van Humbeeck, J.; Kruth, J. Selective laser melting to manufacture “in situ” metal matrix composites: A review. Adv. Eng. Mater. 2019, 21, 1801244. [Google Scholar] [CrossRef]
  36. Yu, W.; Sing, S.; Chua, C.; Kuo, C.; Tian, X. Particle-reinforced metal matrix nanocomposites fabricated by selective laser melting: A state of the art review. Prog. Mater. Sci. 2019, 104, 330–379. [Google Scholar] [CrossRef]
  37. Maurya, P.; Kota, N.; Gibmeier, J.; Wanner, A.; Roy, S. Review on study of internal load transfer in metal matrix composites using diffraction techniques. Mater. Sci. Eng. A 2022, 840, 142973. [Google Scholar] [CrossRef]
  38. Qu, M.; Guo, Q.; Escano, L.I.; Yuan, J.; Hojjatzadeh, S.M.H.; Clark, S.J.; Fezzaa, K.; Sun, T.; Chen, L. Controlling melt flow by nanoparticles to eliminate surface wave induced surface fluctuation. Addit. Manuf. 2022, 59, 130873. [Google Scholar] [CrossRef]
  39. Leung, C.L.A.; Marussi, S.; Atwood, R.C.; Towrie, M.; Withers, P.J.; Lee, P.D. In situ X-ray imaging of defect and molten pool dynamics in laser additive manufacturing. Nat. Commun. 2018, 9, 1355. [Google Scholar] [CrossRef]
  40. Qu, M.; Guo, Q.; Escano, L.I.; Nabaa, A.; Fezzaa, K.; Chen, L. Nanoparticle-enabled increase of energy efficiency during laser metal additive manufacturing. Addit. Manuf. 2022, 60, 103242. [Google Scholar] [CrossRef]
  41. Chowdhury, S.; Yadaiah, N.; Prakash, C.; Ramakrishna, S.; Dixit, S.; Gupta, L.R.; Buddhi, D. Laser powder bed fusion: A state-of-the-art review of the technology, materials, properties & defects, and numerical modelling. J. Mater. Res. Technol. 2022, 20, 2109–2172. [Google Scholar] [CrossRef]
  42. Dejene, N.D.; Lemu, H.G. Current status and challenges of powder bed fusion-based metal additive manufacturing: Literature review. Metals 2023, 13, 424. [Google Scholar] [CrossRef]
  43. Liu, T.; Kinzel, E.C.; Leu, M.C. In-situ lock-in thermographic measurement of powder layer thermal diffusivity and thickness in laser powder bed fusion. Addit. Manuf. 2023, 74, 103726. [Google Scholar] [CrossRef]
  44. Liu, T.; Lough, C.S.; Sehhat, H.; Ren, Y.M.; Christofides, P.D.; Kinzel, E.C.; Leu, M.C. In-situ infrared thermographic inspection for local powder layer thickness measurement in laser powder bed fusion. Addit. Manuf. 2022, 55, 102873. [Google Scholar] [CrossRef]
  45. Vallabh, C.K.P.; Zhao, X. Melt pool temperature measurement and monitoring during laser powder bed fusion based additive manufacturing via single-camera two-wavelength imaging pyrometry (STWIP). J. Manuf. Process. 2022, 79, 486–500. [Google Scholar] [CrossRef]
  46. König, H.-H.; Pettersson, N.H.; Durga, A.; Van Petegem, S.; Grolimund, D.; Chuang, A.C.; Guo, Q.; Chen, L.; Oikonomou, C.; Zhang, F.; et al. Solidification modes during additive manufacturing of steel revealed by high-speed X-ray diffraction. Acta Mater. 2023, 246, 118713. [Google Scholar] [CrossRef]
  47. Liu, J.; Wen, P. Metal vaporization and its influence during laser powder bed fusion process. Mater. Des. 2022, 215, 110505. [Google Scholar] [CrossRef]
  48. Qu, M.; Guo, Q.; Escano, L.I.; Clark, S.J.; Fezzaa, K.; Chen, L. Mitigating keyhole pore formation by nanoparticles during laser powder bed fusion additive manufacturing. Addit. Manuf. Lett. 2022, 3, 100068. [Google Scholar] [CrossRef]
  49. Tourret, D.; Klemm-Toole, J.; Castellanos, A.E.; Rodgers, B.; Becker, G.; Saville, A.; Ellyson, B.; Johnson, C.; Milligan, B.; Copley, J.; et al. Morphological stability of solid-liquid interfaces under additive manufacturing conditions. Acta Mater. 2023, 250, 118858. [Google Scholar] [CrossRef]
  50. Sutton, A.T.; Kriewall, C.S.; Karnati, S.; Leu, M.C.; Newkirk, J.W.; Everhart, W.; Brown, B. Evolution of AISI 304L stainless steel part properties due to powder recycling in laser powder-bed fusion. Addit. Manuf. 2020, 36, 101439. [Google Scholar] [CrossRef]
  51. An, K.; Chen, Y.; Stoica, A.D. VULCAN: A “hammer” for high-temperature materials research. MRS Bull. 2019, 44, 878–885. [Google Scholar] [CrossRef]
  52. An, K. Data reduction and interactive visualization software for event mode neutron diffraction. In ORNL Report, ORNL-TM-2012; Oak Ridge National Laboratory: Oak Ridge, TN, USA, 2012; p. 621. Available online: https://neutrons.ornl.gov/sites/default/files/VDRIVE%20manual_1.pdf (accessed on 1 July 2024).
  53. Toby, B.H. EXPGUI, a graphical user interface forGSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  54. Larson, A.C.; Von Dreele, R.B. General structure analysis system (GSAS). In Los Alamos National Laboratory Report LAUR; Los Alamos National Laboratory: Los Alamos, NM, USA, 2004; pp. 86–748. Available online: https://www.scirp.org/reference/referencespapers?referenceid=105526 (accessed on 1 July 2024).
  55. Shi, N.; Bourke, M.A.M.; Roberts, J.A.; Allison, J.E. Phase-stress partition during uniaxial tensile loading of a TiC-particulate-reinforced Al composite. Metall. Mater. Trans. A 1997, 28, 2741–2753. [Google Scholar] [CrossRef]
  56. Yu, D.; Huang, L.; Chen, Y.; Komolwit, P.; An, K. Real-time in situ neutron diffraction investigation of phase-specific load sharing in a cold-rolled TRIP sheet steel. JOM 2018, 70, 1576–1586. [Google Scholar] [CrossRef]
  57. Chen, Y.; Cernatescu, I.; Venkatesh, V.; Stoica, A.D.; An, K. On the residual stress relaxation in Inconel 718 superalloys at high temperature by real-time neutron diffraction. Mater. Des. 2023, 232, 112135. [Google Scholar] [CrossRef]
  58. Gruber, B.; Weißensteiner, I.; Kremmer, T.; Grabner, F.; Falkinger, G.; Schökel, A.; Spieckermann, F.; Schäublin, R.; Uggowitzer, P.J.; Pogatscher, S. Mechanism of low temperature deformation in aluminium alloys. Mater. Sci. Eng. A 2020, 795, 139935. [Google Scholar] [CrossRef]
Figure 1. In situ neutron diffraction setup. (a) A photo of the experimental setup for the in situ high-temperature compression test (side view). (b) Schematic illustration of the in situ neutron experiment optics at VULCAN (top view). The induction heating coil and the thermocouple are not shown in (b) for simplicity.
Figure 1. In situ neutron diffraction setup. (a) A photo of the experimental setup for the in situ high-temperature compression test (side view). (b) Schematic illustration of the in situ neutron experiment optics at VULCAN (top view). The induction heating coil and the thermocouple are not shown in (b) for simplicity.
Metals 14 01064 g001
Figure 2. Compressive stress–strain curve for Al6061 and Al6061+TiC composites at 250 °C.
Figure 2. Compressive stress–strain curve for Al6061 and Al6061+TiC composites at 250 °C.
Metals 14 01064 g002
Figure 3. In situ neutron diffraction during compressive loading at 250 °C. (a) Diffraction pattern and Rietveld refinement of Al6061+5%TiC before loading at 250 °C. (b) Diffraction pattern evolution of Al6061+5%TiC.
Figure 3. In situ neutron diffraction during compressive loading at 250 °C. (a) Diffraction pattern and Rietveld refinement of Al6061+5%TiC before loading at 250 °C. (b) Diffraction pattern evolution of Al6061+5%TiC.
Metals 14 01064 g003
Figure 4. Chemistry changes of Al6061+5%TiC during loading at 250 °C. (a) Two-dimensional contour plot showing the peak intensity evolution of TiC {200} and Mg2Si {220}. (b) Evolution of the weight fraction of the TiC and Mg2Si phases. (c) Stable chemistry of TiC indicated by the evolution of the modeled Al occupancy at the Ti site.
Figure 4. Chemistry changes of Al6061+5%TiC during loading at 250 °C. (a) Two-dimensional contour plot showing the peak intensity evolution of TiC {200} and Mg2Si {220}. (b) Evolution of the weight fraction of the TiC and Mg2Si phases. (c) Stable chemistry of TiC indicated by the evolution of the modeled Al occupancy at the Ti site.
Metals 14 01064 g004
Figure 5. Lattice strain evolution during compressive loading of Al6061 and Al6061+TiC at 250 °C. (ac) hkl-specific lattice strains determined by SPF. (df) average phase-specific lattice strains determined by Rietveld refinement of the full diffraction pattern, which were then compared with the lattice strain of {311} via SPF. The large fitting error bars for the TiC phase in Al6061+2%TiC stem from the smaller TiC fraction and thus the weaker peak intensity.
Figure 5. Lattice strain evolution during compressive loading of Al6061 and Al6061+TiC at 250 °C. (ac) hkl-specific lattice strains determined by SPF. (df) average phase-specific lattice strains determined by Rietveld refinement of the full diffraction pattern, which were then compared with the lattice strain of {311} via SPF. The large fitting error bars for the TiC phase in Al6061+2%TiC stem from the smaller TiC fraction and thus the weaker peak intensity.
Metals 14 01064 g005
Figure 6. Evolution of “phase stress” (ac), deviatoric stress (df), and “hydrostatic stress” (gi) during compressive loading. The “phase stresses” were calculated on the basis of the lattice strains obtained from the Rietveld refinement.
Figure 6. Evolution of “phase stress” (ac), deviatoric stress (df), and “hydrostatic stress” (gi) during compressive loading. The “phase stresses” were calculated on the basis of the lattice strains obtained from the Rietveld refinement.
Metals 14 01064 g006
Figure 7. Evolution of peak intensity for different Al lattice planes during compressive loading. (ac) Evolution of peak intensity during compression at room temperature. (df) Evolution of peak intensity during compression at 250 °C. RT represents room temperature. The peak intensity during loading was normalized by dividing it by the initial peak intensity before loading for each lattice plane.
Figure 7. Evolution of peak intensity for different Al lattice planes during compressive loading. (ac) Evolution of peak intensity during compression at room temperature. (df) Evolution of peak intensity during compression at 250 °C. RT represents room temperature. The peak intensity during loading was normalized by dividing it by the initial peak intensity before loading for each lattice plane.
Metals 14 01064 g007
Figure 8. Evolution of peak width for different Al lattice planes during compression of (a) Al6061, (b) Al6061+2%TiC, (c) Al6061+5%TiC at 250 °C. The peak width was calculated by FWHM/d. FWHM is the peak width at the half maximum, and d is the lattice spacing. The peak width during loading was normalized by dividing it by the initial peak width before loading for each lattice plane.
Figure 8. Evolution of peak width for different Al lattice planes during compression of (a) Al6061, (b) Al6061+2%TiC, (c) Al6061+5%TiC at 250 °C. The peak width was calculated by FWHM/d. FWHM is the peak width at the half maximum, and d is the lattice spacing. The peak width during loading was normalized by dividing it by the initial peak width before loading for each lattice plane.
Metals 14 01064 g008
Table 1. Chemical composition of as-received Al6061 powders.
Table 1. Chemical composition of as-received Al6061 powders.
ElementAlMgSiCuCrFeTiMnZnOther
Weight percentageBalance0.860.550.270.10.090.01<0.01<0.01<0.15
Table 2. Processing parameters for LPBF process.
Table 2. Processing parameters for LPBF process.
Laser PowerBeam DiameterExposure TimePoint DistanceHatching Spacing
200 W70 µm200 µs80 μm80 μm
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Qu, M.; Yu, D.; Chen, L.; An, K.; Chen, Y. Understanding the High-Temperature Deformation Behaviors in Additively Manufactured Al6061+TiC Composites via In Situ Neutron Diffraction. Metals 2024, 14, 1064. https://doi.org/10.3390/met14091064

AMA Style

Qu M, Yu D, Chen L, An K, Chen Y. Understanding the High-Temperature Deformation Behaviors in Additively Manufactured Al6061+TiC Composites via In Situ Neutron Diffraction. Metals. 2024; 14(9):1064. https://doi.org/10.3390/met14091064

Chicago/Turabian Style

Qu, Minglei, Dunji Yu, Lianyi Chen, Ke An, and Yan Chen. 2024. "Understanding the High-Temperature Deformation Behaviors in Additively Manufactured Al6061+TiC Composites via In Situ Neutron Diffraction" Metals 14, no. 9: 1064. https://doi.org/10.3390/met14091064

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