Home Synthesis, microstructure, and mechanical properties of in situ TiB2/Al-4.5Cu composites
Article Open Access

Synthesis, microstructure, and mechanical properties of in situ TiB2/Al-4.5Cu composites

  • Hongying Li EMAIL logo , Shouxin Zhao , Yangxun Ou and Yongqiu Lai
Published/Copyright: October 18, 2016

Abstract

In situ TiB2/Al-4.5Cu composites with different TiB2 particle amounts were fabricated by the salt-metal reaction technique. The effects of in situ TiB2 on the microstructure and mechanical properties of Al-4.5Cu alloy were studied in this paper. The results showed that in situ TiB2 particles had significant effect on refining grain size and improving mechanical properties of as-cast Al-4.5Cu alloy. With the amounts of TiB2 particles increasing, the yield strength and ultimate tensile strength were improved, while the elongation reduced. The strengthening mechanisms of the in situ particle-reinforcing Al matrix composites were discussed, and the yield strength was predicted accurately by accounting for the three strengthening mechanisms and particle distribution.

1 Introduction

Particulate-reinforced aluminum matrix composites (AMCs) have been widely used in the field of aerospace and automotive industries for its favorable mechanical properties including high elasticity modulus, specific strength, and low coefficient of linear expansion [1], [2], [3]. AMCs have been traditionally fabricated by liquid metallurgy route via ex situ technique [4] or powder metallurgy (P/M) [5]. Both the processing routines are fabricated by directly adding reinforcing particles into molten matrix or powder form, and the particles are prepared separately before the fabrication process of the composites [6]. However, several drawbacks may be produced by these routines. First, it limits the scale of the reinforcing phase, which is always more than 1 μm in size. Second, the composites processed by these methods suffer from interfacial reactions and poor wettability between the reinforcing particles and the matrix, due to surface oxidation and contamination of the reinforced particles [7]. So, uniformly dispersed fine and thermally stable ceramic particulates are very necessary to obtain optimum mechanical properties. The in situ processing techniques were greatly developed to over these shortcomings. The in situ reinforcing particles are usually synthesized through inner chemical reactions between different elements during the composite fabrication. The in situ composites become increasingly attractive for several advantages [8], [9]. First, the endogenously synthesized particles are thermodynamically stable without any surface oxidation and contamination, leading to better wettability and well bonding between the particle and matrix. Second, the in situ particles are finer than the traditional routines. The fine in situ particles refine the grain structure during solidification, resulting in better mechanical properties. During the past decade, in situ TiB2/Al composites have become a promising candidate for traditional AMCs due to the desirable properties of TiB2 particles. TiB2 exhibits many excellent advantages such as high melting point, high hardness, and high stiffness characteristics [6], [10]. The use of K2TiF6 and KBF4 has been well developed to introduce the TiB2 particles. The chemical reactions are as follows [9]:

(1)K2TiF6+AlTiAl3+KAlF4+K3AlF6
(2)KBF4+AlAlB2+KAlF4
(3)AlB2+TiAl3TiB2

In order to facilitate the development of particulate-reinforced AMCs, it is also necessary to develop constitutive relationships that can be used to predict the mechanical properties of AMCs as a function of reinforced particle, matrix, and processing conditions. Investigation about the influence of in situ TiB2 on the microstructure and mechanical properties of the AMCs has been reported numerously in the literatures [9], [10], [11]. However, so far, the strengthening mechanism of the in situ TiB2/Al composite was rarely studied systematically for the complex distribution of the in situ particles. Some modeling works [12], [13], [14] have been done to predict particle-reinforced metal matrix composites over the past few years. Nonetheless, most models are based on the uniform distribution of the reinforced particles. Actually, the distribution of the particles in the matrix has an important influence on the enhancement effect and determines the dominant mechanism [15], [16]. Orowan strengthening, thermal mismatch strengthening, and Hall-Petch strengthening are known as the main strengthening mechanisms of the submicron-sized particle-reinforced AMCs. If the particles are pushed to the grain boundaries, Hall-Petch strengthening will occur resulting from growth restriction-type grain refining. If they are engulfed by the growing grains, the Orowan strengthening and thermal mismatch strengthening will act as the main strengthening mechanisms [15]. In the case of the in situ TiB2/Al composite, it is more complicated for the reason that both the particles in the grain and on the grain boundaries can be attributed to the grain refining. It is of great necessity to consider the particle distribution when predicting the mechanical properties of AMCs as a function of reinforced particle, matrix, and processing conditions.

In this work, TiB2/Al-4.5Cu composites were fabricated by the in situ reaction of K2TiF6 and KBF4 salts in the Al melt. The effect of TiB2 particles on the microstructures and mechanical properties of Al-4.5Cu alloy were studied. The influences of TiB2 particle content and distribution on yield strength of in situ TiB2/Al-4.5Cu composites were discussed via predicted models as the focus.

2 Materials and methods

The TiB2 particle amounts in TiB2/Al-4.5Cu composites were designed as 0 vol.%, 1.7 vol.%, 3.6 vol.%, and 7.2 vol.%, respectively. According to the conversion relationship between volume fraction and mass fraction [17], different amounts of K2TiF6 and KBF4 were weighed and prepared. The preparation process schematic diagram of the in situ TiB2/Al-4.5Cu composite is shown in Figure 1A. In each of these cases, Al-4.5Cu alloy was melted at 850°C in an electric resistance furnace. Then, mixtures of K2TiF6 and KBF4 salts were added to the melt in the atomic ratio in accordance with Ti/2B to favor the formation of TiB2 particles. Prior to the addition of the salt mixture, it was preheated at 300°C for 2 h to get rid of the moisture and volatile impurities. The salt mixture was slowly added into the melt to facilitate the in situ reaction. The melt was gently stirred using a graphite rod every 10 min in a 30-min reaction time. After the reaction, the byproducts, such as the slag containing KAlF4 and K3AlF6, were removed by decanting the crucible. The molten composite was then poured into a preheated (200°C) mild steel mold of size: 150 mm×150 mm×12 mm. Elementary analysis was conducted to calculate the volume fractions of the TiB2 particles from the mass fraction tested.

Figure 1: The preparation process schematic diagram and tensile specimen size of the in situ TiB2/Al-4.5Cu composite. (A) The schematic diagram of preparation process; (B) the tensile specimens size.
Figure 1:

The preparation process schematic diagram and tensile specimen size of the in situ TiB2/Al-4.5Cu composite. (A) The schematic diagram of preparation process; (B) the tensile specimens size.

The phase compositions of composites were identified by A Rigaku D/Max2500 X-ray diffractometer (Rigaku Corporation, Tokyo, Japan) using Cu Kα radiation. The morphology of the composites was characterized using both the optical microscope (Leica DMI300, Leitz, Wetzlar, Germany) and a scanning electron microscope (Sirion 200, FEI Company, Germany). The average grain size of each sample was obtained using the mean linear intercept method on polished samples of alloy and the composites. The tensile property test was carried out on a MTS810 tester at a strain rate of 1×10−2 s−1. The tensile samples were fabricated according to ASTM: E8/E8M-11 standards. Figure 1B shows the tensile specimen size of the in situ TiB2/Al-4.5Cu composite. To ensure the repetitiveness of stress-strain curves, three samples were tested for each composite.

3 Results and discussion

3.1 Microstructure characterization

Figure 2 shows the X-ray diffraction (XRD) patterns of the in situ TiB2/Al-4.5Cu composites with peaks corresponding to the Al, TiB2, and CuAl2. The intensity of TiB2 peaks increases with the increase in the TiB2 content in the composites.

Figure 2: XRD diffraction patterns of the in situ TiB2/Al–4.5Cu composites.
Figure 2:

XRD diffraction patterns of the in situ TiB2/Al–4.5Cu composites.

Figure 3A~B shows the SEM micrographs of the as-cast 1.7% TiB2/Al-4.5Cu composite and 7.2% TiB2/Al–4.5Cu composite, respectively. It can be observed that a mass of continuous CuAl2 phases distribute along the grain boundaries, and the continuous CuAl2 phases transformed into discontinuous phases gradually with the increasing TiB2 content. From the higher magnification microstructure at the top right corner of Figure 3, it can be seen that most of the TiB2 particles distribute along the grain boundaries and interweave with CuAl2 phases together. Figure 3D~E shows the element area scanning by EDS of 7.2% TiB2/Al-4.5Cu composite. Most of the Ti and Cu element are distributed in the grain boundaries, which can further confirm the distribution characteristic of the TiB2 particles. This is different from the previous work that reported that the TiB2 particles could uniformly distribute in α-Al [9], [18]. The primary reason for this is that the ratio of the thermal conductivity coefficient between the particles and matrix is <1. According to the theory of thermal conductivity, during solidification, the TiB2 particles will be easily pushed onto the solidifying interface [19]. Agglomeration of TiB2 particles has also been observed in the composites, which might affect the mechanical properties adversely.

Figure 3: SEM micrographs of TiB2/Al composites. (A) 1.7TiB2/Al-4.5Cu; (B) 7.2TiB2/Al-4.5 Cu, and surface distribution of the element of Al (C), Cu (D), and Ti (E) in 1.7TiB2/Al-4.5Cu composite.
Figure 3:

SEM micrographs of TiB2/Al composites. (A) 1.7TiB2/Al-4.5Cu; (B) 7.2TiB2/Al-4.5 Cu, and surface distribution of the element of Al (C), Cu (D), and Ti (E) in 1.7TiB2/Al-4.5Cu composite.

Figure 4 shows SEM at higher magnification and TEM micrograph of as-cast 7.2.vol.% TiB2/Al-4.5Cu composites, which presents the typical morphology and size distribution of TiB2 particles. The in situ formed TiB2 particles show a mixture of different shapes such as spherical, hexagonal, and cubic, which is in accordance with the previous work in the literatures [19], [20]. It can also be seen in Figure 4B that the interface between the TiB2 particles and the matrix is clean and clear. The well-bonded interface can be attributed to the in situ formation process within the Al melt. Besides, some θ′ phases around the TiB2 particles can be seen. The θ′ phases may be precipitated during the casting cooling process because the distortion field around the TiB2 particles could promote the precipitation of an unstable phase. During the reaction process, the in situ formed TiB2 particles were protected from the oxidation or other contamination. Meanwhile, because of the thermodynamic stability of the TiB2 particle, it avoided producing undesirable compounds between the particle and the matrix during the heat-preservation process [21], [22]. The contamination of reinforcements would lead to the poor wettability between the reinforcements and the matrix [7]. So, the in situ formed uncontaminated particles are good at preferable wettability and interfacial integrity between the aluminum matrix and the TiB2 particle. Figure 4C reveals that the in situ formed TiB2 particles in the composite are between 60 and 400 nm in size. The average particle size dp is 200 nm using the Image pro plus 6.0 software, much lower than that of the ex situ particles in discontinuously reinforced composites [6].

Figure 4: The typical (A) SEM (BSE) micrographs of the TiB2 particle; (B) TEM micrographs of the TiB2 particle; (C) size distribution of TiB2 particles in TiB2/Al-4.5Cu composites at as-cast state.
Figure 4:

The typical (A) SEM (BSE) micrographs of the TiB2 particle; (B) TEM micrographs of the TiB2 particle; (C) size distribution of TiB2 particles in TiB2/Al-4.5Cu composites at as-cast state.

Figure 5 shows the effect of TiB2 content on grain structure and size of as-cast xTiB2/Al-4.5Cu composites with a different particle content (x=0, 1.7, 3.6, and 7.2 vol.%). Figure 5A shows large equiaxed grains with coarsening dendritic substructure and great branch distance. Continuous phases were observed at the grain boundaries. The grain size of the casting base alloy was about 205.0 μm. Figure 5B~D illustrates that average grain size of the as-cast composite decreases with increasing particle content, which means that the TiB2 particles have significant effect on refining grain size of as-cast composites. When the content of the TiB2 particles reaches 7.2 vol.%, all the grains of the as-cast composite are almost fine and equiaxed. Figure 5E shows the relationship between the average grain size of the TiB2/Al-4.5Cu composites and particle content. As the amount of TiB2 particles increases, the grain size reduces. The grain size reaches as low as 40.8 μm with 7.2 vol.% TiB2. Based on the published literature [23], [24], the high grain-refining effect is mainly due to the TiB2 particles nucleating in the composites. However, it should be accepted that any nucleus acting as nucleating agents should be located at the center of the grain. In this work, most TiB2 crystals are pushed into the grain boundaries and distribute in the intergranular region, indicating that they do not nucleate α-Al alone. The main reason for the refined grain size may be attributed to the influence of particles pinning on grain boundaries, which effectively limits grain growth [1], [16].

Figure 5: Effect of TiB2 content on grain structure and size of (A) Al-4.5Cu; (B) 1.7TiB2/Al-4.5Cu; (C) 3.6TiB2/Al-4.5Cu; (D) 7.2TiB2/Al-4.5Cu composites; (E) Shows the average grain size of composites.
Figure 5:

Effect of TiB2 content on grain structure and size of (A) Al-4.5Cu; (B) 1.7TiB2/Al-4.5Cu; (C) 3.6TiB2/Al-4.5Cu; (D) 7.2TiB2/Al-4.5Cu composites; (E) Shows the average grain size of composites.

3.2 Mechanical properties and strengthening mechanisms

Table 1 and Figure 6 present the tensile properties [Young’s modulus, yield strength (R0.2) at 0.2% strain, ultimate tensile strength (Rm), and El% (elongation)] of Al-4.5Cu base alloy and TiB2/Al-4.5Cu composites at room temperature.

Table 1:

Mechanical properties of TiB2/Al-4.5Cu composites and Al-4.5Cu alloy.

SamplesActual TiB2 content (vol.%)Young’s modulusYield strengthElongation (%)UTS (MPa)
Al4.5Cu066.315712.1166
Al4.5Cu/TiB21.771.21647.8199
3.673.71736.4207
7.275.51815.0212
Figure 6: Effect of TiB2 content on the tensile mechanical properties.
Figure 6:

Effect of TiB2 content on the tensile mechanical properties.

In comparison to the base alloy, the in situ composites showed a significant improvement in the yield strength and ultimate tensile strength, while the elongation reduced with the increase in TiB2 particle content. It is accepted that the increase in yield strength by incorporating TiB2 is commonly attributed to the following three mechanisms: grain refinement strengthening [14], [23], Orowan strengthening [12], [25], and coefficient of thermal expansion (CTE) mismatch strengthening [13]. For grain refinement strengthening, the effect of grain size d on the mechanical properties of a composite can be commonly described by the Hall-Petch relationship [26], [27]:

(4)σy=σo+kyd1/2

where, the parameter σ0 represents the yield strength of a single crystal in the absence of any strengthening mechanisms except the solid solution effect, and ky is the magnitude by which crystal boundaries in a polycrystalline material resist slip. In this experiment, taking values of ky=0.04 MPam for Al and Al alloy [28]. In cases where the addition of particles reduces the size of the grains in the particulate-reinforced AMCs, the yield strength improvement due to grain refinement can be described by the following equation [29]:

(5)ΔσHall-Petch=ky(dMMC12dm12)

where, dMMC and dm are the average grain diameters of polycrystalline matrix materials in the composite and unreinforced base alloy. This equation assumes that the Hall-Petch parameters ky and σo remain unchanged in the composites during processing.

Figure 4 illustrates that the TiB2 size is less than 1 μm; so, the Orowan strengthening should not be neglected in estimating the yield strength. The particulate-dislocation interaction by means of the Orowan bowing mechanism in the matrix can increase the strength. The yield strength contributed from the Orowan strengthening is given as in the following equations [13]:

(6)ΔσOrowan=0.13Gmbλlndp2b
(7)Gm=E2(1+ν)
(8)λdp[(12Vp)1/31]

where, Gm is the shear modulus of the matrix; b is the Burgers vector of the dislocations, b=0.286 nm; ν is the Poisson’s ratio, E is the elastic modulus of the matrix, E(Al-4.5Cu)=66.3 Gpa; λ is the interparticle spacing; dp is the particle diameter, dp=200 nm; Vp is the volume fraction of reinforced particles.

Because of the difference in the coefficient of thermal expansion (CTE) and the elastic modulus (Ec) mismatch between the matrix and reinforced-particle particles, enormous dislocations are produced as the result of residual plastic strain. CTE mismatch strengthening relates the contribution of dislocation density to the strength of the material. The effect of thermal mismatch on the strength of the composite is given [12], [13]:

(9)ΔσCTE=βGmbρCTE
(10)ρCTE=12ΔαΔTVpbdp(1Vp)

where β is a constant and approximately equal to 1.25. ρCTE is the dislocation density induced by the CTE mismatch. Δα is the difference between the CTE of the matrix and reinforced particle, αAl=24.6×10−6; αTiB2=7.8×106. ΔT is the difference between the processing and test temperatures, Ttest=298 K; Tprocess=1123 K.

There are a lot of modeling methods to predict the yield strength of particle-reinforced aluminum composites. The arithmetic summation method proposed by relevant scholars [14] is one of the simplest and easily achieved prediction modeling. Obviously, it is assumed that different mechanisms do not influence each other, and they can independently contribute to the final yield strength of composites. According to the arithmetic summation method, the yield strength of the TiB2/Al-Cu composite in this study could be predicted by the following equation [14]:

(11)σy=σm+ΔσHall-Petch+ΔσOrowan+ΔσCTE

Figure 7 shows the experimental and predicted yield strength according to the arithmetic summation method. The parameters used in the predicted calculation of the yield strength are shown in Table 2. The predicted values using the arithmetic summation method are much higher than the experimental values at the as-cast state. The high discrepancies are mainly due to the micrometric agglomerate of TiB2 particulates on the grain boundaries. As seen from the SEM microstructure in Figure 3, the TiB2 particles are not in ideal uniform distribution among the composites. There must be less effective TiB2 particles introducing the strengthening effect than the nominal TiB2 particle content used in the theoretic calculation. Resultantly, the predicted values of yield strength should be overestimated in TiB2/Al composites. In fact, Orowan strengthening, CTE mismatch strengthening, and Hall-Petch strengthening are closely linked to the distribution of the reinforced particles [13]. The Orowan mechanism and CTE mismatch strengthening just act on effective particulates in the grain; the micrometric agglomerate particulates on the grain boundaries should be excluded in the calculation [16]. So, the actual interparticle distance was larger than the theoretical distance λ calculated from equation (8). In particle pushing, reinforced particles are pushed ahead of the solidification front and can act to restrict grain growth. The particles pushed to grain boundaries only can contribute to the grain refinement [15]. So, the fraction in the grain and on the grain boundaries must have a great effect on the strength of the composites.

Figure 7: Theoretically calculated yield strength using the arithmetic summation method.
Figure 7:

Theoretically calculated yield strength using the arithmetic summation method.

Table 2:

Detailed parameters of different TiB2 particles reinforcing Al matrix composites.

CompositesE (matrix)/GPaSizeTemperature (K)CTE (×10−6/K)
Matrix grain/μmParticle (AVG)/nmProcessTestMatrixParticle
TiB2/Al-4.5Cu
 1.7%66.3125.2200112329824.67.8
 3.6%78.7
 7.2%40.8
[30] TiB2/Al-7Si
 0%69.063.01150107329824.67.8
 3%41
 6.5%36
[24] TiB2/Al-4Cu
 1.5%64.931500107329824.67.8
 3%27
 5%21
[8] TiB2/A356
 1.5%72.48.050042829824.67.8
 3%6.0
 5%3.5
[31] TiB2/7055Al
 0%72.626.1275112329824.67.8
 3%16.4
 6.5%10.2

In order to consider the effect of the particle distribution on the yield strength, it is assumed that the proportion of the particle content in grain and on grain boundaries is n/m. Then, the yield strength of the composites can be expressed as equation (12):

(12)σy=σm+ΔσHall-Petch+ΔσOrowan+ΔσCTE=σm+k(dMMC12dm12)+0.13Gmbdp[(12nn+mVp)131]lndp2b+kGmb12nn+mVpΔαΔT(1nn+mVp)bdp

Figure 8A shows the comparison between the yield strength prediction using a different ratio of n:m with the experimental data in this study. As we can see, the yield strength prediction of the composites decreases with the decrease in the ratio of n to m experimental data. When all the particles distribute in the grain, the yield strength reaches a maximum value, the main strengthening mechanisms are Orowan strengthening and CTE mismatch strengthening. When n/m is zero, all the particles distribute on the grain boundaries, the strengthening effect is lowest, only the Hall-Petch effect contributes to the yield strength. The multistrengthening mechanisms work when the particles are distributed both in the grain and on the grain boundaries. The strengthening effect weakens as the proportion of particles inside the grain decreases. Compared to the experimental value in the study, it can be found that the predicted values agree well with the experimental value when n/m is 1/50. It means that the proportion of the particle content in grain and on grain boundaries is 1/50, which confirms the same conclusion in Figure 3 that most of the TiB2 particles distribute along the grain boundaries.

Figure 8: The comparison of yield strength prediction by the modified model with experimental data for TiB2/Al alloy composites. (A) in this study; (B) deriving from reference.
Figure 8:

The comparison of yield strength prediction by the modified model with experimental data for TiB2/Al alloy composites. (A) in this study; (B) deriving from reference.

Figure 8B shows a comparison between the predicted results using the appropriate ratio of n:m and the experiment results of yield strength for TiB2/Al composites deriving from reference, which further verify the accuracy of the modified method. In these literatures, several types of in situ TiB2/Al alloy composites are included, such as TiB2/Al-7Si [30] composites, TiB2/Al-4Cu [24] composites, TiB2/A356 composites [8], and TiB2/7055Al composites [31]. It is easy to see that the predicted results show good agreement with the experimental values by choosing the suitable value of the proportion of the particle content on grain boundaries and in grain. Generally, it is indicated that the phenomenological model can give a fairly precise estimation for the yield stress of the in situ TiB2/Al alloy composites.

4 Conclusions

  1. The sub-micro-sized TiB2 particle is produced by the salt-metal reaction technique using K2TiF6 and KBF4. The addition of TiB2 can refine the gain size of cast Al-4.5Cu alloy and reach a significant improvement in the yield strength and ultimate tensile strength of cast Al-4.5Cu alloy, while the elongation reduces. When TiB2 particles reached 7.2 vol.%, the yield strength and ultimate tensile strength of TiB2/Al-4.5Cu composites reached 181 MPa and 212 MPa, respectively; however, the elongation was reduced to 5%.

  2. The main strengthening mechanisms of the in situ TiB2/Al composites include grain refinement strengthening, Orowan strengthening, and coefficient of thermal expansion mismatch strengthening. Big discrepancies existed in predicting the yield strength of in situ TiB2/Al composites when making arithmetic summation of the three strengthening mechanisms without considering the agglomerates of TiB2 particles on the grain boundaries. The arithmetic summation method was modified by considering the distribution of particles and introducing the ratio n:m of the particle content in grain and on grain boundaries. When n:m was chose as 1:50, the predicted values of the yield strength agreed well with the experimental value.


Corresponding author: Hongying Li, School of Materials Science and Engineering, Central South University, Changsha 410083, Hunan, People’s Republic of China; and Nonferrous Metal Oriented Advanced Structural Materials and Manufacturing Cooperative Innovation Center, Central South University, Changsha 410083, China, e-mail:

Acknowledgments

This work was supported by the Changsha City Science and Technology Program (no. K1303013-11).

References

[1] Zhang Z, Topping T, Li Y, Vogt R, Zhou Y, Haines C. Scripta Mater. 2011, 65, 652–655.10.1016/j.scriptamat.2011.06.037Search in Google Scholar

[2] Deng KK, Wang XJ, Wu YW, Hu XS, Wu K, Gan WM. Mater. Sci. Eng. A. 2012, 543, 158–163.10.1016/j.msea.2012.02.064Search in Google Scholar

[3] Tian SF, Jiang LT, Guo Q, Wu GH. Mater. Des. 2014, 53, 129–136.10.1016/j.matdes.2013.06.038Search in Google Scholar

[4] Hao S. Sci. Eng. Compo. Mater. 2015, 22, 673–684.10.1515/secm-2013-0184Search in Google Scholar

[5] Lakshmipathy J. Sci. Eng. Compo. Mater. 2015, 22, 573–582.10.1515/secm-2013-0268Search in Google Scholar

[6] Lü L, Lai MO, Su Y, Teo HL, Feng CF. Scripta. Mater. 2001, 45, 1017–1023.10.1016/S1359-6462(01)01128-9Search in Google Scholar

[7] Tjong SC, Ma ZY. Mater. Sci. Eng. R. 2000, 29, 49–113.10.1016/S0927-796X(00)00024-3Search in Google Scholar

[8] Mandal A, Chakraborty M, Murty BS. Mater. Sci. Eng. A. 2008, 489, 220–226.10.1016/j.msea.2008.01.042Search in Google Scholar

[9] Lu L, Lai MO, Chen FL. Acta Mater. 1997, 45, 4297–4309.10.1016/S1359-6454(97)00075-XSearch in Google Scholar

[10] Yi H, Ma N, Zhang Y, Li X, Wang H. Scripta Mater. 2006, 54, 1093–1097.10.1016/j.scriptamat.2005.11.070Search in Google Scholar

[11] Wang M, Chen D, Chen Z, Wu Y, Wang F, Ma N. Mater. Sci. Eng. A. 2014, 590, 246–254.10.1016/j.msea.2013.10.021Search in Google Scholar

[12] Zhang Z, Chen DL. Mater. Sci. Eng. A. 2008, 483, 148–152.10.1016/j.msea.2006.10.184Search in Google Scholar

[13] Zhang Z, Chen DL. Scripta Mater. 2006, 54, 1321–1326.10.1016/j.scriptamat.2005.12.017Search in Google Scholar

[14] Sanaty-Zadeh A. Mater. Sci. Eng. A. 2012, 531, 112–118.10.1016/j.msea.2011.10.043Search in Google Scholar

[15] Schultz BF, Ferguson JB, Rohatgi PK. Mater. Sci. Eng. A. 2011, 530, 87–97.10.1016/j.msea.2011.09.042Search in Google Scholar

[16] Kang YC, Chan SLI. Mater. Chem. Phys. 2004, 85, 438–443.10.1016/j.matchemphys.2004.02.002Search in Google Scholar

[17] Mu EZ, Li Q. Hot Working Technology 2013, 42, 115–116.10.1016/j.fuproc.2013.03.045Search in Google Scholar

[18] Yi HZ, Mang NH, Li XF, Zhang YJ, Wang HW. Mater. Sci. Eng. A. 2006, 419, 12–17.10.1016/j.msea.2005.10.020Search in Google Scholar

[19] Han Y, Liu X, Bian X. Compos. A: Appl. Sci. Manufact. 2002, 33, 439–444.10.1016/S1359-835X(01)00124-5Search in Google Scholar

[20] Lakshmi S, Lu L, Gupta M. J. Mater. Process. Technol. 1998, 73, 160–166.10.1016/S0924-0136(97)00225-2Search in Google Scholar

[21] Ashok KB, Murugan N. Mater. Des. 2012, 40, 52–58.10.1016/j.matdes.2012.03.038Search in Google Scholar

[22] Davidson AM, Regener D. Compos. Sci. Technol. 2000, 60, 865–869.10.1016/S0266-3538(99)00151-7Search in Google Scholar

[23] Michael Rajan HB, Ramabalan S, Dinaharan I, Vijay SJ. Mater. Des. 2013, 44, 438–445.10.1016/j.matdes.2012.08.008Search in Google Scholar

[24] Mandal A, Maiti R, Chakraborty M, Murty BS. Mater. Sci. Eng. A. 2004, 386, 296–300.10.1016/S0921-5093(04)00938-4Search in Google Scholar

[25] Lloyd DJ. Int. Mater. Rev. 1994, 39, 1.10.1179/imr.1994.39.1.1Search in Google Scholar

[26] Ibrahim IA, Mohamed FA, Lavernia EJJ. Mater. Sci. 1991, 26, 1137–1156.10.1007/BF00544448Search in Google Scholar

[27] Li Y, Zhao YH, Ortalan V, Liu W, Zhang ZH, Vogt RG. Mater. Sci. Eng. A. 2009, 527, 305–316.10.1016/j.msea.2009.07.067Search in Google Scholar

[28] Hansen N. Scripta Mater. 2004, 51, 801–806.10.1016/j.scriptamat.2004.06.002Search in Google Scholar

[29] Wang T, Chen Z, Zheng Y, Zhao Y, Kang H, Gao L. Mater. Sci. Eng. A. 2014, 605, 22–32.10.1016/j.msea.2014.03.021Search in Google Scholar

[30] Kumar S, Chakraborty M, Subramanya Sarma V, Murty BS. Wear 2008, 265, 134–1342.10.1016/j.wear.2007.09.007Search in Google Scholar

[31] Zhong LH, Zhao YT, Zhang SL, Chen G, Chen S, Liu YH. Trans. Nonferrous Met. Soc. China 2013, 23, 2502–2508.10.1016/S1003-6326(13)62761-2Search in Google Scholar

Received: 2015-11-20
Accepted: 2016-8-27
Published Online: 2016-10-18
Published in Print: 2018-4-25

©2018 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Articles in the same Issue

  1. Frontmatter
  2. Original articles
  3. Review of the mechanical performance of variable stiffness design fiber-reinforced composites
  4. Exact solution for bending analysis of functionally graded micro-plates based on strain gradient theory
  5. Synthesis, microstructure, and mechanical properties of in situ TiB2/Al-4.5Cu composites
  6. Microstructure and properties of W-Cu/1Cr18Ni9 steel brazed joint with different Ni-based filler metals
  7. Drilling studies on the prepared aluminum metal matrix composite from wet grinder stone dust particles
  8. Studies on mechanical properties of thermoplastic composites prepared from flax-polypropylene needle punched nonwovens
  9. Design of and with thin-ply non-crimp fabric as building blocks for composites
  10. Effect of coir fiber reinforcement on mechanical properties of vulcanized natural rubber composites
  11. Investigation and analysis of glass fabric/PVC composite laminates processing parameters
  12. Abrasive wear behavior of silane treated nanoalumina filled dental composite under food slurry and distilled water condition
  13. Finite element study into the effects of fiber orientations and stacking sequence on drilling induced delamination in CFRP/Al stack
  14. Preparation of PAA/WO3 composite films with enhanced electrochromism via layer-by-layer method
  15. Effect of alkali treatment on hair fiber as reinforcement of HDPE composites: mechanical properties and water absorption behavior
  16. Integration of nano-Al with one-step synthesis of MoO3 nanobelts to realize high exothermic nanothermite
  17. A time-of-flight revising approach to improve the image quality of Lamb wave tomography for the detection of defects in composite panels
  18. The simulation of the warpage rule of the thin-walled part of polypropylene composite based on the coupling effect of mold deformation and injection molding process
  19. Novel preparation method and the characterization of polyurethane-acrylate/ nano-SiO2 emulsions
  20. Microwave properties of natural rubber based composites containing carbon black-magnetite hybrid fillers
  21. Simulation on impact response of FMLs: effect of fiber stacking sequence, thickness, and incident angle
Downloaded on 13.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/secm-2015-0478/html
Scroll to top button