Effect of chromium concentration on microstructure and properties of Fe–3.5B alloy
Shengqiang Ma,Jiandong Xing,Guofeng Liu,Dawei Yi,Hanguang Fu,Jianjun Zhang,Yefei Li
DOI: https://doi.org/10.1016/j.msea.2010.07.066
2010-01-01
Abstract:The cast low carbon Fe–3.5B alloys containing various chromium concentrations were prepared in a 10 kg medium frequency induction furnace and the effects of chromium concentration on microstructure and properties of Fe–3.5B alloys have been examined by means of optical microscope (OM), scanning electron microscope (SEM), back-scattered electron microscope (BSE), electron probe microanalyzer (EPMA), energy dispersive spectrum (EDS), X-ray diffraction (XRD), transmission electron microscopy (TEM) and Vickers hardness. As a result, the as-cast structures of Fe–3.5B– X Cr ( X = 0, 2, 5, 8, 12, 18, mass fraction) alloys are mainly composed of dendrite ferrite, martensite, pearlite and boride. The boride in the alloy without chromium addition comprises the eutectic Fe 2 B, which is continuous netlike or fish-bone structure distributed over the metallic matrix. With the increase of chromium concentration in Fe–3.5B alloy, matrix structure turns into the supersaturated α -Fe solid solution while the morphology of boride becomes dispersed due to the transformation of boride from simple Fe 2 B to (Fe,Cr) 2 B when the chromium concentration in Fe–3.5B alloy exceeds 8 wt.%. Meanwhile, some primary M 2 B-type borides may precipitate under this condition. The bulk hardness of the as-cast alloy ranges from 41.8 to 46.8 HRC. However, the bulk hardness of the heat treated alloy rises first and falls later mainly because of the morphology variation of structure. Fracture toughness of boride is improved gradually owing to the entrance of chromium into Fe 2 B, which may be attributed to the change of spatial structure of boride. Keywords Fe–B alloy (Fe,Cr) 2 B boride Fracture toughness Solid solution 1 Introduction Considerable amount of economic loss is brought because of abrasion and corrosion in mechanical parts of machinery and equipment. Therefore, in order to reduce this loss, the research on superior wear-resistant and corrosion-resistant materials has been allotted a high priority in the field of material science recently [1–3] . It is of very important significance to develop excellent wear- and corrosion-resistant materials so as to meet the rigorous environment requirements. The invention of high chromium white cast iron was considered a breakthrough because of its increased toughness compared with plain white cast iron and Ni-hard white cast iron, which is attributed to the improvement of carbide morphology [4,5] . However, the high chromium white cast iron is still a kind of brittle material that cannot meet the requirement of demanding working conditions. Hence, constant research activities are being carried out to produce advanced materials, which are much better than existing wear-resistant materials [6,7] . Recently, much attention has been paid to high boron white cast iron owing to the unique characteristics of boron in steel [8–10] . Investigations have discovered that hardenability, toughness and wear resistance can be improved by adding boron as an alloying element. Since the borides exist in the high boron white cast iron, the study on high boron white cast iron has been widely emphasized by researchers [11–13] . Boride has a higher hardness than carbide when combined with the same element, which gives us the idea to replace the carbides in white cast iron with borides [14] . In addition, carbon is almost undissolved in boride, thus the properties of matrix can be adjusted by carbon content. This should in turn result in improving ductility and fracture toughness, while maintaining the inherently excellent wear resistance. The significance of this approach is that the matrix and hard phases can be adjusted independently by varying carbon and boron contents [4,15,16] . However, there exist many interconnect net-like borides in solidification structure of high boron white cast iron, which damages the continuity of matrix, reduces the toughness and results in fracture in impact environment. Thus, the application of high boron cast alloy is restricted to some extent. Heat treatment and hot deformation are the most common methods used to improve the toughness of cast alloys [17–20] . Nevertheless, heat treatment has little effect on morphology of boride because of lower boron solubility in high boron white cast iron. And hot deformation not only increases the production process and energy consumption, but also is only suitable for the simple-shaped work pieces. Guo and Wang [21,22] reported the solubility and diffusion of boron in Fe–Cr–B alloy. Unfortunately, there is little research about the influence of chromium concentration on the microstructure and properties of Fe–B alloy. The objective of present work aims to study the microstructures and properties of borides in Fe–3.5B alloys with different amount of chromium. As a result, the morphology and microstructure of the boride are improved and the fracture toughness is increased accordingly. 2 Experimental procedure 2.1 Preparation for Fe–3.5B cast alloy The investigated alloys in present work are hypoeutectic high boron white cast iron. The alloys were prepared in a 10 kg medium frequency induction furnace. Initial charge materials were clean low silicon pig iron and steel scrap. Ferro-alloys such as Fe–63 wt.% Cr and Fe–16.28 wt.% B were added to a slag-free molten alloy so as to minimize the oxidation loss and slag formation. When all the alloys were melted in the furnace, 0.10 wt.% pure aluminum was added into the molten alloy to deoxidize at the temperature of 1520–1540 °C. The melt was subsequently super-heated to 1560 °C and transferred into a pre-heated teapot ladle. After removal of any dross and slag, the melt was poured at 1430 °C into the sodium silicate–CO 2 bonded sand moulds, obtaining Y-block ingots following ASTM A781/A 781-M95. The test specimens were cut from the lower part of the Y-block and surface ground to remove 3 mm from the surface and eliminate any oxidized layer. Some of the as-cast specimens were treated at 980 °C for 1 h, quenched in water and then tempered at 200 °C, followed by air cooling. The chemical compositions of Fe–3.5B alloys are listed in Table 1 . 2.2 Microstructure examination The microanalysis of the specimens was carried out using a scanning electron microscope (SEM), a back-scattered electron image (BSE), an X-ray diffraction (XRD), an electron probe microanalysis (EPMA), a transmission electron microscopy (TEM) and an energy dispersive spectrum (EDS) to identify microstructures. For metallographic observations, all specimens were etched in a 4 vol.% nital solution. XRD was performed on a MXP21VAHF diffractometer with Copper Kα radiation coupling continuous scanning at 40 kV and 200 mA as an X-ray source. The specimen was scanned in the 2 θ ranging from 20° to 100° with scanning speed of 2°/min and step space of 0.02°. 2.3 Vickers microindentation fracture toughness test The bulk hardness was measured on an HR-150A Rockwell-hardness tester. The microhardness of boride in Fe–3.5B alloy was also measured by using an HXD-type 1000 Vickers-hardness tester with a load of 100 gf, according the ASTM E384 standard. The indentation cracking method proposed by Palmqvist in 1957 [23] was utilized and the equation used for calculating fracture toughness was as follows [24–29] : (1) K IC = X ⋅ P c 3 / 2 where X is the residual indentation coefficient, which depends on hardness to modulus ratio ( H / E ) of the boride. The constant X is 0.064 ( E / H ) 1/2 , where E and H are the Young's modulus and microhardness, respectively. P is the applied load and the value of E is approximately 290 GPa [24–30] for fracture toughness calculation, and c is the indentation radial cracking length as shown in Fig. 1 . 3 Results 3.1 Solidification microstructure of Fe–3.5B alloys with various chromium concentrations The as-cast solidification microstructures of Fe–3.5B alloys containing various chromium concentrations are shown in Fig. 2 . Fig. 3 is the XRD analysis results of as-cast specimens. From Fig. 2 , the as-cast Fe–3.5B alloys with various chromium concentrations have formed the metallic matrix and eutectic morphologies during the solidification. According to the XRD results (shown in Fig. 3 ), the microstructure of Fe–3.5B alloy is mainly composed of ferrite, martensite, a number of pearlite and eutectic Fe 2 B, which is continuously distributed over the metallic matrix (specimen C0 shown in Fig. 2 (a)). As the chromium concentration increases, there is a tendency that the eutectic boride seems to be scattered. Namely, the continuous netlike or fish-bone structure reduces slightly (shown in Fig. 2 (b)–(d)). When the chromium concentration exceeds 8 wt.%, the morphology of microstructure becomes significantly different. The characteristics that the eutectic boride is continuously distributed over the metallic matrix vanish rapidly. Instead, a great deal of dispersed eutectic boride appears (as shown in Fig. 2 (e) and (f)). Moreover, some primary borides precipitate in the matrix (shown in Fig. 2 (e) and (f)). XRD patterns (C4 and C5 specimen shown in Fig. 3 ) also indicate the presence of many M 2 B-type eutectic borides (M represents both Fe and Cr) in the as-cast Fe–3.5B alloy. SEM analysis of Fe–3.5B alloys with different chromium concentration is shown in Fig. 4 , and Fig. 4 (a) and (c) show the SEM images of as-cast specimen C3 and C5, respectively. The results from EPMA lines in Fig. 4 (a) and (c) indicate that the iron content is lower in boride than in matrix while the chromium is dominant in boride (shown in Fig. 4 (b) and (d)). 3.2 Heat treated microstructure of Fe–3.5B alloy with various chromium concentrations Fig. 5 shows the heat treated microstructures of Fe–3.5B alloy with various chromium concentrations (specimen C2 and C5). From SEM images of specimen C2 and C5 ( Fig. 5 (a) and (b)), it can be seen that there is no difference in eutectic boride between as-cast and heat-treated specimens. However, a number of secondary particles precipitate from the matrix when the quenched specimens (980 °C for 1 h, quenched in water) are tempered at 200 °C (as shown in Fig. 5 (a)–(d)). In addition, more secondary particles precipitate in the matrix of C5 specimen in comparison with that of C2 specimen because of the higher chromium additions (BSE images in Fig. 5 (c) and (d)). Cross-sectional elemental distribution of Fe and Cr by X-ray mapping in Fig. 5 also reveals that the Cr element prevails in the boride (shown in Fig. 5 (e) and (f)). From Fig. 5 , it also indicates that the matrix of Fe–3.5B with high chromium concentration is supersaturated α -Fe solid solution after quenched in water, which leads to the precipitation of many borocarbides. Hence, it can be concluded from the precipitation of secondary particles that the martensite of Fe–3.5B alloys with boron, carbon and chromium is supersaturated solid solution. 3.3 Hardness and fracture toughness of boride in Fe–3.5B alloy with different chromium concentrations The bulk hardness of as-cast and heat treated specimens for C0–C5 alloys are shown in Fig. 6 . Fig. 7 shows the microhardness of matrix in as-cast and heat treated C0–C5 specimens. From Fig. 6 and Fig. 7 , it can be seen clearly that bulk hardness and the microhardness of matrix in as-cast alloys increase with the increase of chromium concentration while heat treated specimens for C0–C5 alloys exhibit different. As the chromium concentration increases, the bulk hardness of heat treated specimens raises firstly, and then falls when the chromium concentration is up to 8 wt.%. This is mainly attributed to morphology change of netlike borides as well as the solution of chromium in the metallic matrix when chromium concentration is increased to certain extent. The chromium concentration in boride was measured by EPMA analysis and the relationships between the fracture toughness and Vickers hardness and chromium concentration in boride of Fe–3.5B alloys are shown in Fig. 8 . It is clearly observed from Fig. 8 , that the Vickers hardness of boride has changed sharply when the chromium element is added to the Fe–3.5B alloys. As chromium content is increased gradually, in comparison with no addition of chromium, Vickers hardness ascends continuously. Its value ranges from 1356.9 to 1859.8 HV 0.1 , namely, about 13.6–18.6 GPa, which is mainly ascribed to the addition of chromium and the formation of Cr-rich boride. As discussed below, Fe atoms in Fe 2 B are substituted by Cr atoms, thus leading to the formation of (Fe,Cr) 2 B. Dybkov [31] also found that the microhardness values vary considerably and the microhardness of (Fe,Cr) 2 B in boronizing layer, HV 0.1 , is 18.0 ± 1.0 GPa, which is also attributed to the non-homogeneity of boride in the boronizing layers. 4 Discussion 4.1 Effect of chromium on microstructure of Fe–3.5B alloy The chromium concentration has a remarkable effect on the as-cast microstructure of Fe–3.5B alloy (as shown in Fig. 2 ). Fig. 9 is the EDS analysis of Fe–3.5B alloys with different chromium concentration in Fig. 4 . Comparing EDS spectrum results of point 1 and point 2 in Fig. 4 (a) (EDS spectrum results shown in Fig. 9 (a) and (b)), the concentration of carbon and silicon elements in matrix of Fe–3.5B alloy is relatively high and these elements promote the formation of pearlite [32,33] . So there exists certain amount of pearlite in the as-cast matrix (as shown in Fig. 2 ). Fig. 9 (c) and (d) is the EDS spectrum of point 3 and point 4 in Fig. 4 (c). Seen from Fig. 9 (a)–(d), it can be obviously observed that the concentration of chromium in borides varies sharply when the chromium is added into Fe–3.5B alloy. Fig. 10 shows the distribution of chromium concentration in matrix and boride measured by EPMA analysis when various chromium concentrations are added into Fe–3.5B alloys. With the increase of chromium concentration in Fe–3.5B alloys, the differential value of chromium concentration between boride and matrix remarkably increases, especially when the chromium concentration in Fe–3.5B alloy exceeds 8 wt.%. It indicates that most of chromium exists in boride. As the chromium concentration increases, the structure of boride is changed (detected by XRD in Fig. 3 ). When the chromium concentration exceeds 12 wt.%, i.e., 18 wt.% for C5 specimen, the atom ratio of Fe to Cr is approximately equal to 1:1 (shown in Fig. 9 (d)), the borides are likely transformed from tetragonal Fe 2 B to orthorhombic (Fe,Cr) 2 B [34–36] . Hence, according to EDS and XRD results, compositionally, the boride may consist of (Fe,Cr) 2 B and Fe 2 B as the chromium concentration exceeds 12 wt.% [37,38] . That is to say, the phase transformation behavior of the alloy would be affected by the presence of high chromium additions [34,35] . The investigations of Ref. [39–41] also confirmed that the Fe–Cr–B alloy coatings, containing the borides of Cr 2 B, Cr 1.65 Fe 0.35 B 0.96 and α -(Fe,Cr) solid solution were formed. However, X-ray diffraction analysis practically does not show the presence of Cr 2 B phase, probably because, firstly, the M 2 B phase and Cr 2 B phase have very similar crystal structures and, secondly, under non-equilibrium conditions, the alloy containing high chromium appears to be single-phase structurally and two-phase compositionally [35,36] . As the chromium concentration is up to 18 wt.% (shown in Fig. 4 (c)), it can be seen that more M 2 B (M represents both Fe and Cr) primary borides precipitate and look like lath distributed in the matrix. Interestingly, there exists a concentration gradient of chromium element near the primary boride (shown in Fig. 4 (c) grey regions adjacent to lath boride). This is probably attributed to the first solidification of primary boride occurs under normal solidification conditions, which consumes more chromium elements, resulting in the formation of chromium depleted zone. Fig. 11 is the bright field TEM micrographs and corresponding selected area diffraction patterns of boride in specimen C0 and C5, respectively. From the as-cast specimen C0 with no chromium addition, shown in Fig. 11 (a) (fish-bone structure of Fe 2 B phase in dark region), it can be obtained from electron diffraction patterns, that the dark region in Fig. 11 (a) can be indexed to Fe 2 B phase, which is a body-centered tetragonal structure and its lattice constant is: a = b = 0.51093 nm, c = 0.42486 nm (namely, c / a = 0.83), whereas M 2 B-type boride in specimen C5 (lath structure of primary M 2 B-type boride shown in dark region of Fig. 11 (b)) is orthorhombic structure. The results analyzed above can also prove the phase transformation of boride from Fe 2 B to (Fe,Cr) 2 B with the increase of high chromium concentration. 4.2 Effect of chromium on composition of M 2 B-type boride precipitated in the Fe–3.5B alloy According to the liquid phase surface of Fe–15%Cr–C–B quaternary system phase diagram [37,42] , the temperature of the investigated Fe–Cr–B alloy, which is adjacent to the precipitation starting temperature of primary borides, is about 1490 °C. It can be seen that M 2 B primary boride generates from the liquid. Taking into consideration the fact that the subsequent solidification of Fe–Cr–B alloy is more complex and it suffers under non-equilibrium solidification condition, thus the primary boride should also precipitate from the melt. From the SEM observation in Fig. 5 , it can be seen that heat treatment has little effects on morphology of boride (shown in Fig. 5 (a) and (b)) mainly owing to the small boron solubility of 0.02 wt.% in austenite [4,21] . Also, primary borides are distributed in the form of long rod in the matrix, and the eutectic borides adjacent to primary boride show a more dispersed distribution while high chromium concentration is added (such as C5 specimen shown in Fig. 5 (b)–(d)). However, more secondary particles precipitate from the matrix (BSE images shown in Fig. 5 (c) and (d)) during the heat treatment when the chromium element is added into Fe–3.5B alloy. XRD result of specimen C5 after heat treatment in Fig. 12 indicates that the secondary precipitation is M 23 (C,B) 6 phase (M presents Fe and Cr) [21] . Moreover, the element surface distribution patterns of Cr and Fe indicate that the Cr element prevails in the M 2 B-type boride (shown in Fig. 5 (e) and (f)). This is because the atom radius ( R Cr = 0.185 nm, R Fe = 0.0.172 nm) and electronegativity ( X P Cr = 1.66 , X P Fe = 1.83 ) of Cr and Fe are very similar, Cr element can dissolve in iron to form a continuously substitutional solid solution, and most Cr atoms replace the atoms of Fe in Fe 2 B lattice. Therefore, it is possible that the replacement of iron atoms undergoes constantly, which leads to the formation of M 2 B boride. So, during the solidification, Cr atoms as the substitution of partial Fe atoms enter the boride. The investigations of Ref. [9,43–45] also indicate the distribution of alloying elements is not homogeneous and boron mainly exists in grain boundaries, only a small amount of boron is distributed in the matrix. So the M 2 (B,C) borocarbide has formed in the B-rich region at the grain boundary, and the chromium element is mainly distributed over the M 2 (B,C) borocarbide. The average composition of eutectic boride for as-cast C0–C5 specimens is confirmed by EPMA measurement, and the EDS results are shown in Table 2 . Table 3 [34,35] is also the composition results of the phases by EPMA measurement in boride layers at Fe–25%Cr alloy–boron interface. Compared with the results in Table 3 , the boride in Fe–3.5B alloy with high chromium concentration mainly consists of (Fe,Cr) 2 B phase, which also proves the presence of (Fe,Cr) 2 B phase. 4.3 Effect of chromium on fracture toughness of boride in Fe–3.5B alloy The fracture toughness ( K IC ) of boride in Fe–3.5B alloys, estimated by the Palmqvist method, displays great difference as the chromium concentration is increased. When the chromium concentration is less than 10.56 wt.% in boride (namely, chromium concentration of 5 wt.% in Fe–3.5B alloy as shown in Fig. 10 ), K IC value (15.3–16.56 MPa m 1/2 as shown in Fig. 8 ) improves slightly, whereas it enhances rapidly from 16.56 to 22.4 MPa m 1/2 (shown in Fig. 8 ) as the chromium concentration ranges from 8.28 to 36.64 wt.% (namely, chromium concentration of 5–18 wt.% in Fe–3.5B alloy as shown in Fig. 10 ). This is because partial Cr element dissolves into the metallic matrix, which reduces the amount of Cr element entering into the boride. When the chromium concentration in Fe–3.5B alloy exceeds 8 wt.%, a large amount of Fe 2 B is transformed into (Fe,Cr) 2 B phase. As discussed above, Fe 2 B has a bbc tetragonal, C16, CuAl 2 -type structure, and its lattice parameters are a = 0.5109 nm, c = 0.4249 nm, c / a = 0.832, while (Fe,Cr) 2 B is orthorhombic structure [37,46] , which may result in the increase of K IC value. Moreover, bond energy of Fe 2 B phase between Fe and B atom is not distributed uniformly at different crystal orientation throughout space, and atomic bond of B–B in Fe 2 B phase is very weak in [0 0 2] direction [47] , which is likely to cause brittleness of Fe 2 B phase, and the addition of chromium may enhance bonding energy of Fe–Fe and B–B atomic bond in weaker direction on spatial distribution and improve brittleness of boride in Fe–3.5B alloy. So the properties of boride are improved. 5 Conclusions The morphology and distribution of borides in Fe–3.5B alloys as well as fracture toughness of borides are investigated and the present work supports the following conclusions: (1) The structure of as-cast Fe–3.5B– X Cr ( X = 0, 2, 5, 8, 12, 18, mass fraction) alloys is mainly composed of dendrite ferrite, martensite, pearlite and borides. Boride without chromium addition comprises Fe 2 B, which is continuous netlike or fish-bone structure distributed over the metallic matrix. (2) With the increase of chromium concentration in Fe–3.5B alloy, matrix structure turns into supersaturated α -Fe solid solution while the morphology of boride becomes dispersed, which is attributed to the transformation of boride from simple Fe 2 B to (Fe,Cr) 2 B when the chromium concentration in Fe–3.5B alloy exceeds 8 wt.%. Meanwhile, some primary M 2 B-type borides may occur under this condition. (3) As chromium concentration is increased, bulk hardness of the as-cast alloy ranges from 41.8 to 46.8 HRC while the heat treated alloy rises firstly and falls later mainly because of the morphology change of the boride. (4) Fracture toughness of boride is improved gradually owing to the entrance of chromium atoms into the Fe 2 B, which may be attributed to the change of spatial structure of boride. Acknowledgements The authors would like to appreciate Prof. J. D. Xing and Dr. H. G. Fu for fruitful suggestions and also thank the financial support for this work from the National High-Tech R & D program under Contract NO. 2007AA03Z510 , P. R. China. References [1] T.C. Zhang D.Y. Li Mater. Sci. Eng. A 325 2002 87 97 [2] X. Wang D.Y. Li Mater. Sci. Eng. A 315 2001 158 165 [3] R. Liu D.Y. Li Wear 225–229 1999 968 974 [4] Z.L. Liu Y.X. Li X. Chen K.H. Hu Mater. Sci. Eng. A 486 2008 112 116 [5] W. Fairhurst K. Rohrig Foundry Trade J. 136 1974 685 698 [6] Z.H. Hu D.C. Luan Foundry Technol. 27 2006 153 155 [7] S. Seo S. Matsuoka T. Kanayama A. Yorifuji J. Iron Steel Inst. Jpn. 89 2003 593 600 [8] H.G. Fu Q. Xiao J.C. Kuang Z.Q. Jiang J.D. Xing Mater. Sci. Eng. A 466 2007 160 165 [9] H.G. Fu Z.H. Li J.Q. Jiang J.D. Xing Mater. Lett. 61 2007 4504 4507 [10] H.G. Fu Z.H. Li Y.P. Lei Z.Q. Jiang J.D. Xing Mater. Des. 30 2009 885 891 [11] K. Peev M. Radulovic M. Fiset J. Mater. Sci. Lett. 13 1994 112 114 [12] M. Radulovic M. Fiset K. Peev M. Tomovic J. Mater. Sci. 29 1994 5085 5094 [13] M. Fiset K. Peev M. Radulovic J. Mater. Sci. Lett. 12 1993 615 617 [14] C.Q. Guo C.D. Wang X.P. Liu P.M. Kelly China Foundry 5 2008 28 31 [15] E. Ya. Gol’dshtein V.G. Mizin Met. Sci. Heat Treat. 30 1988 479 484 [16] H. Baker ASM Handbook: Alloy Phase Diagram [M] tenth ed. 1990 ASM, International Handbook Committee Materials Park, OH [17] A.S. Chaus F.I. Rudnitskii Met. Sci. Heat Treat. 31 1989 121 128 [18] A.S. Chaus Met. Sci. Heat Treat. 47 2005 53 61 [19] A.R. Khodabandeh M. Jahazi S. Yue S.T. Aghdashi Mater. Manuf. Process. 21 2006 105 110 [20] A.R. Khodabandeh M. Jahazi S. Yue P. Bocher ISIJ Int. 45 2005 272 280 [21] C.Q. Guo P.M. Kelly Mater. Sci. Eng. A 352 2003 40 45 [22] W.D. Wang S.H. Zhang X.L. He Acta Metall. 43 1995 1693 1699 [23] S.G. Cook J.E. King J.A. Little Mater. Sci. Technol. 11 1995 1093 1098 [24] A.H. Ucisik C. Bindal Surf. Coat. Technol. 94–95 1997 561 565 [25] U. Sen S. Sen S. Koksal F. Yilmaz Mater. Des. 26 2005 175 179 [26] S. Sen I. Ozbek U. Sen C. Bindal Surf. Coat. Technol. 135 2001 173 177 [27] B.R. Lawn E.R. Fuller J. Mater. Sci. 10 1997 2016 2024 [28] B.A. Cook A.M. Russell J.L. Harringa A.J. Slager M.T. Rohe J. Alloys Compd. 366 2004 145 151 [29] I. Campos R. Rosas U. Figueroa C. VillaVelázquez A. Meneses A. Guevara Mater. Sci. Eng. A 488 2008 562 568 [30] N. Frantzevich F.F. Voronov S.A. Bakuta Elastic Constants and Elastic Modulus for Metals and Nonmetals: Handbook 1982 Naukova Dumka Press Kiev [31] V.I. Dybkov J. Mater. Sci. 42 2007 6614 6627 [32] X.H. Zhi J.D. Xing Y.M. Gao Mater. Sci. Eng. A 487 2008 171 179 [33] R. Benz J.F. Elliott J. Chipman Metall. Mater. Trans. B 5 1974 2235 2240 [34] V.I. Dybkov W. Lengauer K. Barmak J. Alloys Compd. 398 2005 113 122 [35] V.I. Dybkov W. Lengauer P. Gas J. Mater. Sci. 41 2006 4948 4960 [36] M. Carbucicchio G. Palmobarini M. Rateo G. Sambogna Hyperfine Interact. 116 1998 143 148 [37] P. Christodoulou N. Calos Mater. Sci. Eng. A 301 2001 103 117 [38] V.I. Dybkov L.V. Goncharuk V.G. Khoruzha K.A. Meleshevich A.V. Samelyuk V.R. Sidorko Solid State Phenom. 138 2008 181 188 [39] H.W. Jin C.G. Park M.C. Kim Mater. Sci. Eng. A 304–306 2001 321 326 [40] H.W. Jin Y.M. Rhyim S.G. Hong C.G. Park Mater. Sci. Eng. A 304–306 2001 1069 1074 [41] H.W. Jin C.G. Park M.C. Kim Surf. Coat. Technol. 113 1999 103 112 [42] N.J. Calos E. Graham D.R. Cousens P. Christodoulou C.H.L. Kennard L.K. Bekessy Mater. Trans. 42 2001 496 501 [43] S.H. Zhang X.L. He T. Ko J. Mater. Sci. 29 1994 2655 2662 [44] M. Jahazi J.J. Jonas Mater. Sci. Eng. A 335 2002 49 61 [45] X. Huang M.C. Chaturvedi N.L. Richards J. Jackman Acta Mater. 45 1997 3095 3107 [46] C. Bindal A. Hikmet Ucisik Surf. Coat. Technol. 122 1999 208 213 [47] M.S. Li S.L. Fu W.D. Xu R.L. Zhang R.H. Yu Acta Metall. Sin. 31 1995 201 207