Improved wear resistance of Al–15Si alloy with a high current pulsed electron beam treatment
Y. Hao,B. Gao,G.F. Tu,S.W. Li,C. Dong,Z.G. Zhang
DOI: https://doi.org/10.1016/j.nimb.2011.04.010
2011-01-01
Abstract:A hypereutectic Al–15Si alloy (Si 15 wt.%, Al balance) was irradiated by high current pulsed electron beam (HCPEB). The HCPEB treatment causes ultra-rapid heating, melting and cooling at the top surface layer. As a result, the special “halo” microstructure centering on the primary Si phase is formed on the surface due to interdiffusion of Al and Si elements. The composition of the “halo” microstructure is distributed continuously from the center to the edge of the “halo”. Compared to an untreated matrix, the remelted layer underneath the surface presents single contrast because of the compositional homogeneity after HCPEB treatment. The thickness of the remelted layer increases slightly from 4.4 μm (5 pulses) to 5.6 μm (25 pulses). HCPEB treatment broadens and shifts the diffraction peaks of Al and Si. The lattice parameters of Al decreases due to the formation of a supersaturated solid solution of Al in the melted layer. Through analysis of Raman spectra and transmission electron microscopy (TEM), the amorphous Si (a-Si) and nanocrystalline Si are formed in the near-surface region under multiple bombardments of HCPEB. The relative wear resistance of a 15-pulse sample is effectively improved by a factor of 9, which can be attributed to the formation of metastable structures. Keywords Hypereutectic Al–15Si alloy High current pulsed electron beam “Halo” microstructure Supersaturated solid solution Amorphous Si Nanocrystalline Si 1 Introduction Hypereutectic Al–Si alloys (>12 wt.% Si) are composed of the primary Si phase and the eutectic matrix [1] . Because of their low thermal expansion coefficient, high thermal stability and good wear and corrosion resistance, the hypereutectic Al–Si alloys are considered to be excellent materials for manufacturing engine pistons and rotors used in the automobile industry [2] . However, the coarse primary Si and acicular eutectic Si in the microstructure causes cast hypereutectic Al–Si alloys to have poor properties, for instance, poor machining and ductility. Therefore, it is necessary to modify eutectic and primary Si, especially coarse primary Si, to improve the mechanical properties of these materials. In past decades, intense pulsed energetic beams, such as laser, ion and electron beams, have been applied to modify the surfaces of materials [3–8] . Among these pulsed beam techniques, high current pulsed electron beam (HCPEB) has quickly developed as a new, high-power energetic beam for surface modification of materials because of its high efficiency, simplicity and reliability [9–13] . HCPEB irradiation induces coupled dynamic temperature and stress fields on the surface of metallic materials, giving rise to superfast heating, melting and cooling processes with rates between 10 7 and 10 9 K/s [9] together with thermal stress and shock waves. The material surfaces experience the nonequilibrium processes induced by the HCPEB treatment. As a result, metastable phases [14,15] are formed in the modified surface layer, which is the origin of significantly enhanced physicochemical and mechanical properties [16–20] that are often unattainable with conventional surface treatment techniques. Some researchers [8,21] have discovered that HCPEB treatment can effectively improve the microhardness and wear resistance of a pure Al surface. The surface carbonization of pure Al through HCPEB irradiation has also been studied. But to our knowledge, there have been few investigations on microstructure and property changes of the top surfaces of HCPEB-treated hypereutectic Al–Si alloys. Therefore, in this paper, a detailed study on the wear property of an Al–15Si alloy surface before and after HCPEB modification is reported for the first time. 2 Experimental procedures 2.1 Experimental equipment and sample preparation The electron beam system used in our experiments is a “Nadezhda-2” HCPEB source [22] . It produces an electron beam with an electron energy between 10 and 40 keV, a short pulse duration between 0.5 and 5 μs, an energy density between 0.5 and 5 J/cm 2 , a peak current density between 10 2 and 10 3 A/cm 2 , a cross-section area between 10 and 50 cm 2 and a repetition rate of 0.1 Hz. The electron pulse is generated by an explosive emission graphite cathode. The beam energy density is controlled by the accelerating voltage, magnetic field intensity and anode-collector distance. The whole HCPEB treatment process is carried out in a vacuum, pressured at 5.5 × 10 −3 Pa. More details about the principles of HCPEB systems are available [8] . Hypereutectic aluminum silicon alloy (Al–15Si) with the following composition is studied in this work: Si 15 wt.% and Al balance. Prior to HCPEB treatment, the samples for microstructural analysis were machined to cylinders with diameters of 10 mm and heights of 9 mm. The sample surfaces were ground with sandpapers of P800, P1000 and P2000 and then polished with 1 μm diamond paste. However, the samples used for wear tests were machined to cylinders with diameters of 6 mm and heights of 13 mm. Then, the sample surfaces were ground and polished until the sample reached a height of 12 mm. Finally, all processing surfaces were cleaned with absolute ethyl alcohol before HCPEB treatment and then dried with a hair drier. The parameters used in the HCPEB surface treatment were as follows: an accelerating voltage of 23 kV, an energy density of 2.5 J/cm 2 , 5, 15 and 25 pulses, and an anode-target distance of 10 cm. 2.2 Microstructure characterization and wear test Surface and cross-sectional morphologies of the samples were examined with SSX-550 scanning electron microscopy (SEM). An XRD-PW3040/60 X-ray diffractometer with a Cu Kα radiation source was used to detect the phase transformation in the surface layer; the scanning range (2 θ ) was from 20° to 100° with a step length of 0.02°. A 1610-type electron probe micro-analysis (EPMA) was performed to observe composition distribution of the melted surface. The structural change of the Si phase was analyzed by HR800 Raman spectra. A TECNAI20 transmission electron microscopy (TEM) was also carried out to further investigate the top surface melted layer. The sample sheets with diameters of 3 mm that were used in the TEM experiment were obtained by mechanical thinning (∼80 μm), dimpling (∼10 μm) and ion beam (Ar + ) thinning (tens of nanometers) of the sample back surfaces. Wear tests under dry conditions were conducted with a MG-2000 pin-on-disc machine at room temperature (25 °C). The untreated and HCPEB-treated samples that had 6-mm diameters and 12-mm lengths were prepared for the wear tests. Stainless steel (1Cr18Ni9) was used as the sliding counterpart disc in all tests. The surface hardness and roughness of stainless steel were 192 HV and 1 μm ( R a ), respectively. The applied load was 10 N. The sliding speed and distance were maintained at 0.8 m/s and 0.38 km, respectively. The weight loss was measured by an electronic balance (±0.1 mg). 3 Results and discussion 3.1 Microstructure characteristics Fig. 1 shows typical surface morphology micrographs of an Al–15Si alloy before and after HCPEB treatment. Fig. 1 (a) presents the microstructure of the initial sample, which consists of a coarse primary Si phase and an Al matrix. Fig. 1 (b)–(d) shows the surface morphologies of HCPEB-treated samples with different numbers of pulses (5, 15 and 25). It can be seen that a special “halo” microstructure centering on the primary Si phase is formed on all irradiated surfaces. The phase boundary of primary Si phase becomes less faceted than that of an untreated sample. In general, the “halo” microstructure of hypereutectic Al–Si alloys is formed by α (Al) growing around the primary Si phase [23,24] . However, in our case, the formation of a “halo” microstructure is a result of interdiffusion of Al and Si elements that surrounds the primary Si phase, which is quite different from the formation mechanism of a “halo” reported in other references. With HCPEB pulses (5 or 15 pulses), the “halo” microstructure centering on the primary Si phase emerges on the top melted surface, as seen in Fig. 1 (b) and (c). When the number of pulses increases (25 pulses), the “halo” microstructure becomes larger and rounder due to diffusion of Si atoms in the liquid state, but some smaller primary Si phases disappear and dissolve into the Al matrix completely, as shown in Fig. 1 (d). As a result, a supersaturated solid solution of Al is formed in the near-surface layer. The chemical composition distribution of the treated surface becomes more homogeneous when the contrast changes of back scattered electron images ( Fig. 1 (a)–(d)) before and after HCPEB modification are compared. Moreover, after HCPEB treatment, cracks appear in the center of the primary Si phase due to the great thermal stresses induced by HCPEB [12,25] . The formation of a “halo” microstructure centering on the primary Si phase can be explained with the HCPEB-induced rapid heating and cooling process (∼10 7 –10 9 K/s) under a high energy density [26,27] . During the HCPEB process, interdiffusion of Al and Si elements occurs around the primary Si phase in the melting state. And then, due to subsequent super-rapid solidification, there is not enough time for the Al and Si elements to redistribute and be back to their original states [28] . Consequently, the special “halo” microstructure is obtained on the top treated surface. A supersaturated solid solution of Al is formed because Si atoms dissolve into the Al matrix. 3.2 Analysis of “halo” microstructure Taking a 15-pulse treated sample as an example, we discuss in detail the typical “halo” microstructure. To further understand the diffusion behavior of Al and Si elements during the HCPEB process, the composition distribution of a “halo” microstructure is analyzed by EPMA. Fig. 2 shows an EPMA image and the composition distribution of the “halo” microstructure after HCPEB treatment with 15 pulses. It is found that the formation of a “halo” microstructure is a result of the interdiffusion of Al and Si elements. Fig. 2 (a) shows a typical morphology of a “halo”. Fig. 2 (b) and (c) shows the area distribution of Al and Si elements. It can be observed that Al and Si elements centering on the primary Si phase are distributed continuously from the center to the edge of the “halo”. Moreover, Si atoms gradually diffuse into the Al matrix, which is observed with the color change around the primary Si phase in Fig. 2 (c). The diffusion behavior of Al atoms is similar to that of Si, as shown in Fig. 2 (b). To summarize, interdiffusion of Al and Si elements in the liquid state plays an important role in the formation of the “halo” microstructure, which can promote fusion of a coarse primary Si phase and an Al matrix. 3.3 Cross-section analysis by SEM Fig. 3 shows cross-sectional SEM morphologies of an Al–15Si alloy after HCPEB treatment. Compared with the matrix, the remelted layers with different microstructure characteristics can be observed, as shown in Fig. 3 . After HCPEB treatment, a remelted layer is formed on the top surface of the modified sample, which presents single contrast due to its compositional homogeneity. It can be seen that there is no evident distinction between the Al and Si phases, and a supersaturated solid solution of Al is formed after HCPEB irradiation. Furthermore, the thickness of the remelted layer increases slightly as the number of HCPEB pulses increases. For 5 pulses ( Fig. 3 (a)), the thickness of the remelted layer is about 4.4 μm, whereas the thickness reaches about 5.6 μm after 25 pulses, as shown in Fig. 3 (b). The phenomenon is also observed with HCPEB-treated DZ4 and steel samples [29,30] . The effect of heat accumulation in underlying substrate materials is considered to be a reasonable explanation for the increasing thickness of the remelted layer with multiple short interval pulses. 3.4 XRD analysis To further analyze the phase transformation of treated surface and lattice distortion of Al lattice, XRD was employed. Fig. 4 shows XRD patterns of an Al–15Si alloy surface before and after HCPEB treatment. Compared with the initial sample, no new phase is found after HCPEB irradiation, as shown in Fig. 4 (a). However, HCPEB treatment broadens the diffraction peaks, which becomes increasingly obvious as the number of pulses increases, which is clearly seen in Fig. 4 (b). HCPEB bombardment also shifts the diffraction peaks of Al and Si toward higher angles (right shift). The largest shift of diffraction peaks appears in the XRD pattern of the 15-pulse sample. The peak broadening and shift may be due to the combined effect of grain refinement, residual stresses [29] and dislocations induced by HCPEB. According to Bragg’s equation, it is found that interplanar spacing d values of Al and Si lattices decreases after HCPEB treatment, as shown in Fig. 5 . In the figure, the d (111) values of Al and Si lattices decrease rapidly and reach minimum values after a 15-pulse treatment, whereas the d (111) values increase again after 25 pulses of HCPEB. The variation of d (111) values is consistent with the analysis results for the diffraction peak shift. In addition, the lattice parameters of the Al lattice were calculated through celref3 [31,32] , and the results are given in Fig. 6 . Compared with those of an untreated sample, the lattice parameters of Al decrease after HCPEB treatment. The value for 15-pulse treated sample decreases from an initial value of 0.40577 to 0.40515 nm. This result can be explained as follows: (1) a supersaturated solid solution of Al is formed in the remelted layer, and (2) Al atoms in the lattice are replaced by Si atoms with small atomic radium, which causes lattice distortion and reduces the Al lattice parameters. However, the Al lattice parameters increase after 25 pulses because of the release of residual stress. 3.5 Raman spectrum analysis The Raman spectra of silicon are clearly different in the crystalline and amorphous states. Crystalline silicon (c-Si) has a strong sharp peak near 520 cm −1 in the Raman spectrum. However, amorphous silicon (a-Si) has a continuous wave packet in the vicinity of 480 cm −1 in the Raman spectrum [33] . Thus, the variation of Si structure before and after HCPEB treatment can be analyzed with the Raman spectra of Si. Fig. 7 shows the Raman spectra of Si on an Al–15Si alloy surface before and after HCPEB treatment, which demonstrates the structural transformation of Si from a crystalline state to an amorphous state. In the initial state, Raman peaks of untreated c-Si appear at 523, 306 and 959 cm −1 , but the spectrum of c-Si is dominated by an intense, narrow peak at 523 cm −1 . The peak at 523 cm −1 in the Raman spectra shows that the Si is in a crystalline state. After HCPEB treatment of an Al–15Si alloy surface, Raman peaks of Si in the Raman spectra change significantly. The intensity of all three c-Si peaks is seen to decrease and finally disappear as the number of HCPEB pulses increases, and a-Si band can be observed for a number of pulses. The position of the a-Si band in the Raman spectra is consistent with that reported in previous results [33] . Compared with an untreated sample, the c-Si 523 cm −1 peak broadens and slightly moves toward low Raman shift after 5 pulses. The spectrum obtained for 15 pulses consists of an a-Si band centered at 468 cm −1 , whereas an a-Si band appears on the Raman shift of 478 cm −1 after 25 pulses. Both bands are considered to be Raman bands of a-Si [33,34] . The characteristic peaks of c-Si at 959 and 306 cm −1 decrease in intensity and eventually disappear, and the sharp peak of c-Si at 523 cm −1 disappears, too. The conversion from a sharp c-Si peak to a broad a-Si band can be attributed to a breakdown in the selection rules of Raman that all phonon modes become active in a-Si [34–36] . According to the above analysis, we conclude that the surface crystalline layer of the Al–15Si alloy is damaged after HCPEB treatment. Si can exist in an amorphous state in some zones of the irradiated surface, which is the result of rapid solidification induced by HCPEB. However, the a-Si phase has not been found in XRD analysis because the content of amorphous Si in the modified layer is small and cannot be detected by XRD equipment [17] . 3.6 TEM analysis of the surface layers Fig. 8 shows a typical bright field TEM image taken from the top surface layer of the 15-pulse treated Al–15Si alloy and the corresponding selected area electron diffraction (SAED) patterns. It can be observed that nanocrystallines are formed and distributed evenly on the top of modified layer after HCPEB treatment of 15 pulses, as shown in Fig. 8 (a). These nanocrystallines are identified as Si by the corresponding SAED pattern in Fig. 8 (b). The result (grain refinement) is consistent with the broadening of diffraction peaks in XRD patterns. The formation of dispersive distribution of nanocrystalline Si is attributed to the mechanism of dissolution and re-precipitation [37] . The rapid heating and cooling induced by HCPEB gives rise to the recrystallization process on the treated surface, and nanocrystalline Si particles precipitate on the top surface during the process. However, an amorphous structure is formed on the treated surface by observing the corresponding SAED pattern, as shown in Fig. 8 (c). There is a possible link between amorphous structure and nanostructure [38] . The formation of an amorphous structure may be attributed to the size of some nanostructure particles below a critical size in local regions [14,17,38] . The result of TEM further proves HCPEB treatment can induce the formation of metastable structures and is coincident with results of SEM (“halo” microstructure) and Raman analysis (a-Si). 3.7 Wear behavior The weight loss of untreated and HCPEB-treated samples was measured before and after the wear tests. Fig. 9 shows the variations in the weight loss of untreated and HCPEB-treated Al–15Si alloy before and after the wear tests. The weight loss is reduced from 1.8 × 10 −3 g of the untreated sample to 0.2 × 10 −3 g of the 15-pulse sample, showing a significant improvement of the sliding wear resistance of Al–15Si alloy after HCPEB treatment. Compared with the initial sample, the weight loss of the sample declines by 88.9% after 15 pulses, and relative wear resistance is improved by a factor of 9. The results can be explained as follows: (1) the formation of a supersaturated solid solution of Al is the main reason for enhanced wear resistance; (2) the presence of nanocrystalline Si and amorphous Si in the surface layer has a hardening effect on wear property, and (3) the dislocations and residual stresses formed in the microstructure can improve wear resistance. However, the weight loss of the 25-pulse sample increases again for two possible reasons: stress release and the second phase may not be as strong. Thus, the wear resistance of the 25-pulse Al–15Si alloy may increase. 4 Conclusions This work has examined the effect of HCPEB treatment on modifying a hypereutectic Al–15Si alloy surface to investigate microstructural change and wear resistance. The main conclusions are summarized as follows: (1) A special “halo” microstructure is formed on all treated surfaces because of the interdiffusion of Al and Si elements surrounding the primary Si phase. The chemical composition of the “halo” microstructure centered on the primary Si phase is distributed continuously from the center to the edge of the “halo”. (2) The remelted layer presents single contrast because of the compositional homogeneity, and a supersaturated solid solution of Al is formed in the near-surface layer after HCPEB treatment. The thickness of the remelted layers increases slightly from 4.4 μm for 5-pulse sample to 5.6 μm for 25-pulse sample. (3) After HCPEB treatment, the diffraction peaks of Al and Si broaden and shift because of the combined effect of grain refinement, dislocations and residual stresses. The lattice parameters of Al decreases due to the formation of a supersaturated solid solution of Al. (4) Raman analysis shows that the surface crystalline layer of the Al–15Si alloy is damaged seriously after HCPEB irradiation. The Raman spectra obtained for 15 and 25 pulses are made up of a-Si bands near 468 and 478 cm −1 . Si can exist in an amorphous state in some zones of the irradiated surface. By TEM observation, nanocrystalline Si particles are distributed uniformly in the top surface layer after 15-pulse treatment. (5) The best wear resistance of Al–15Si alloy surface is obtained for the 15-pulse treatment, and the relative wear resistance is improved by a factor of 9. Acknowledgements The present work is supported by the Fundamental Research Funds for the Central Universities ( N090602009 & N100402010 ), the Education Department of Liaoning Province of China ( 2008T239 & LT2010036 ), and the Key Projects in the National Science & Technology Pillar Program during the Eleventh Five-Year Plan Period ( 2009BAE80B01 ). References [1] J. Chang I. Moon C. Choi Refinement of cast microstructure of hypereutectic Al–Si alloys through the addition of rare earth metals J. Mater. Sci. 33 1998 5015 5023 [2] M. Gupta E.J. Lavernia Effect of processing on the microstructural variation and heat-treatment response of a hypereutectic Al–Si alloy J. Mater. Process Technol. 54 1995 261 270 [3] J. Stasic M. Trtica B. Gakovic Surface modifications of AISI 1045 steel created by high intensity 1064 and 532 nm picosecond Nd:YAG laser pulses Appl. Surf. Sci. 255 2009 4474 4478 [4] A.D. Pogrebnjak V.T. Shablya N.V. Sviridenko Study of deformation states in metals exposed to intense-pulsed-ion beams (IPIB) Surf. Coat. Technol. 111 1999 46 50 [5] X. Wang M.K. Lei J.S. Zhang Surface modification of 316L stainless steel with high-intensity pulsed ion beams Acta Metall. Sin. 201 2007 5884 5890 [6] V. Engelko B. Yatsenko G. Mueller Pulsed electron beam facility (GESA) for surface treatment of materials Vacuum 62 2001 211 216 [7] A.P. Surzhikov T.S. Frangulyan S.A. Ghyngazov Structural-phase transformations in near-surface layers of alumina–zirconium ceramics induced by low-energy high-current electron beams Nucl. Instrum. Methods Phys. Res. Section B 267 2009 1072 1076 [8] C. Dong A. Wu S. Hao Surface treatment by high current pulsed electron beam Surf. Coat. Technol. 163/164 2003 620 624 [9] S.Z. Hao P.S. Wu J.X. Zou Microstructure evolution occurring in the modified surface of 316L stainless steel under high current pulsed electron beam treatment Appl. Surf. Sci. 253 2007 5349 5354 [10] G. Mueller V. Engelko A. Weisenburger Surface alloying by pulsed intense electron beams Vacuum 77 2005 469 474 [11] F.J. Xu G.Z. Tang G.W. Guo Influence of irradiation number of high current pulsed electron beam on the structure and properties of M50 steel Nucl. Instrum. Methods Phys. Res. Section B 268 2010 2395 2399 [12] J.X. Zou K.M. Zhang T. Grosdidier Orientation-dependent deformation on 316L stainless steel induced by high-current pulsed electron beam irradiation Mater. Sci. Eng. A 483–484 2008 302 305 [13] T. Grosdidier J.X. Zou N. Stein Texture modification, grain refinement and improved hardness/corrosion balance of a FeAl alloy by pulsed electron beam surface treatment in the ‘‘heating mode” Scripta Mater. 58 2008 1058 1061 [14] S.Z. Hao Y. Qin X.X. Mei Fundamentals and applications of material modification by intense pulsed beams Surf. Coat. Technol. 201 2007 8588 8595 [15] T. Grosdidier J.X. Zou B. Bolle Grain refinement, hardening and metastable phase formation by high current pulsed electron beam (HCPEB) treatment under heating and melting modes J. Alloy. Compd. 504S 2010 S508 S511 [16] J. An X.X. Shen Y. Lu Microstructure and tribological properties of Al–Pb alloy modified by high current pulsed electron beam Wear 261 2006 208 215 [17] Q.F. Guan H. Zou G.T. Zou Surface nanostructure and amorphous state of a low carbon steel induced by high-current pulses electron beam Surf. Coat. Technol. 196 2005 145 149 [18] B. Gao S.Z. Hao J.X. Zou High current pulsed electron beam treatment of AZ31 Mg alloy J. Vac. Sci. Technol. A 23 2005 1548 1553 [19] B. Gao Y. Hao G.F. Tu Compounded surface modification of ZK60 Mg alloy by highcurrent pulsed electron beam + micro-plasma oxidation Plasma Sci. Technol. 12 2010 67 70 [20] B. Gao S.Z. Hao J.X. Zou Effect of high current pulsed electron beam treatment on surface microstructure and wear and corrosion resistance of an AZ91HP magnesium alloy Surf. Coat. Technol. 201 2007 6297 6303 [21] S.Z. Hao S. Yao J. Guan Surface treatment of aluminum by high current pulsed electron beam Curr. Appl. Phys. 1 2001 203 208 [22] D.I. Proskurovsky V.P. Rotshtein G.E. Ozur Pulsed electron-beam technology for surface modification of metallic materials J. Vac. Sci. Technol. A 16 1998 2480 2488 [23] F. Yilmaz R. Elliott Halo formation in Al–Si alloys Met. Sci. 18 1984 362 366 [24] R.S. Barclay P. Niessen H.W. Kerr Halo formation during unidirectional solidification of off-eutectic binary alloys J. Cryst. Growth 20 1973 175 182 [25] Y. Qin C. Dong Z.F. Song Deep modification of materials by thermal stress wave generated by irradiation of high-current pulsed electron beams J. Vac. Sci. Technol. A 27 2009 430 435 [26] J.X. Zou T. Grosdidier K.M. Zhang Microstructures and phase formations in the surface layer of an AISI D2 steel treated with pulsed electron beam J. Alloy. Compd. 434–435 2007 707 709 [27] S.Z. Hao B. Gao A.M. Wu Surface modification of steels and magnesium alloy by high current pulsed electron beam Nucl. Instrum. Methods Phys. Res. Section B 24 2005 646 652 [28] G. Chen, H.Z. Fu, New Metal Materials of Non-equilibrium Solidification, first ed., Science Press, Beijing, 2004. (in Chinese). [29] S.Z. Hao X.D. Zhang X.X. Mei Surface treatment of DZ4 directionally solidified nickel-based superalloy by high current pulsed electron beam Mater. Lett. 62 2008 414 417 [30] J.X. Zou T. Grosdidier K.M. Zhang Cross-sectional analysis of the graded microstructure in an AISI D2-steel treated with low energy high-current pulsed electron beam Appl. Surf. Sci. 255 2009 4758 4764 [31] H. Abdelkefi H. Khemakhem G. Vélu Dielectric properties of ferroelectric ceramics derived from the system BaTiO 3 –NaNbO 3 -based solid solutions Solid State Sci. 6 2004 1347 1351 [32] M. Balonis F.P. Glasser The density of cement phases Cem. Concr. Res. 39 2009 733 739 [33] Z.Y. Zhu, R.Y. Gu, T.H. Lu, Application of Raman Spectrum in the Chemistry, first ed., Northeastern Press, Shenyang, 1998 (in Chinese). [34] T.A. Harriman D.A. Lucca J.K. Lee Ion implantation effects in single crystal Si investigated by Raman spectroscopy Nucl. Instrum. Methods Phys. Res. Section B 267 2009 1232 1234 [35] M.H. Brodsky M. Cardona Local order as determined by electronic and vibrational spectroscopy: amorphous semiconductors J. Non-Cryst. Solids 31 1978 81 108 [36] T. Hochbauer A. Misra M. Nastasi Physical mechanisms behind the ion-cut in hydrogen implanted silicon J. Appl. Phys. 92 2002 2335 2342 [37] A. Dehghan-Manshadi A. Shokouhi P.D. Hodgson Effect of isothermal dissolution and re-precipitation of austenite on the mechanical properties of duplex structures Mater. Sci. Eng. A 527 2010 6765 6770 [38] H. Gleiter Nanostructured materials: basic concepts and microstructure Acta Mater. 48 2000 1 29