Вы находитесь на странице: 1из 10

Journal of Materials Processing Technology 213 (2013) 453462

Contents lists available at SciVerse ScienceDirect

Journal of Materials Processing Technology


journal homepage: www.elsevier.com/locate/jmatprotec

Friction welding of ductile iron with stainless steel


Radosaw Winiczenko a, , Mieczysaw Kaczorowski b,1
a b

Warsaw University of Life Sciences, Department of Production Engineering, Nowoursynowska 166, 02-787 Warsaw, Poland Warsaw University of Technology, Institute of Mechanics and Design, Narbutta 85, 02-524 Warsaw, Poland

a r t i c l e

i n f o

a b s t r a c t
The study of mechanical properties and microstructure of friction welded coupe of ductile iron with stainless steel are presented. Scanning electron microscopy (SEM) was used for investigation of the fracture morphology and phase transformations taking place during friction welding process. It was concluded that in case of bainitic ductile iron (BDI) the fracture precedes mainly trough the cleavage planes. Moreover, the distribution of selected elements on both side of the joining interface was studied using EDS line and maps spectrometry. The EDS spectrometry showed some enrichment of ductile iron with Cr and Ni atoms close to the joint. The depth of Cr atoms penetration reached 50 m. The heat generated locally by friction increased the temperature in the area close to the interface even over the melting point of ductile iron. This was conrmed by metallography which revealed the carbide eutectic enriched with Cr in ductile iron. 2012 Elsevier B.V. All rights reserved.

Article history: Received 1 April 2012 Received in revised form 9 October 2012 Accepted 17 October 2012 Available online 26 October 2012 Keywords: Friction welding Ductile iron Stainless steel Microstructure

1. Introduction Friction welding is a solid-state joining process which produces coalescence in materials, using the heat generated between surfaces through the combination of a mechanically induced rubbing motion and the applied load. The resulting joints are of forged quality. Under normal conditions, the faying surfaces do not melt. Filler metal, ux and shielding gas are not required with this process. As a rule, all metallic engineering materials that are forgeable can be friction welded, including automotive valve alloys, tool steel, tantalum, alloy steels and maraging steel. In addition, many castings, powder metals and metal matrix composites are weldable. For example, engine pistons, used mainly for truck applications, can be friction welded. In general, there are two different types of material combinationsthe rst being a combination of steel and aluminum, and the second consisting of ductile cast iron with mild steel. Friction welding is applied for welding low ductility materials because it causes crystal renement, the pattern of the heat ow is simple and high compressive residual stresses are involved on the surface of these joints (Lancaster, 1987). According to the American Welding Society (AWS), friction welding is the ideal method for joining metals that are not necessarily similar. Many authors have recently conducted extensive

Corresponding author. Tel.: +48 225 934 624; fax: +48 225 225 934 618/225 934 611. E-mail addresses: radoslaw winiczenko@sggw.pl, rwinicze@poczta.onet.pl (R. Winiczenko), m.kaczorowski@wip.pw.edu.pl (M. Kaczorowski). 1 Tel.: +48 228 494 280; fax: +48 228 483 379. 0924-0136/$ see front matter 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jmatprotec.2012.10.008

investigation into the friction welding of dissimilar materials. The main reasons for dissimilar joining are due to the combination of good mechanical properties of one material and the low specic weight, good corrosion resistance, and good electrical properties of a second material. During the friction welding of dissimilar materials, signicant cost saving is possible because engineers can design bimetallic parts that use expensive materials only where needed. Expensive forgings and castings can be replaced with less expensive forgings to bar steel, tubes, plates and suchlike. Sahin Ahmet et al. (1998) analyzed the friction welding process in relation to the welding of copper to steel bars and subsequently obtained good quality welding joints. Satyanarayana et al. (2005) successfully achieved the joining of austeniticferritic stainless steels using the friction welding process. Sahin (2005) joined highspeed steel (HSS-S 6-5-2) and medium carbon steel (AISI 1040). The strengths of the joints were determined by tension, fatigue and notch-impact tests, and his results were compared with the tensile strength of the base materials. The same author also conducted studies on the friction welding of stainless steel and copper materials (Sahin, 2009). Meshram et al. (2007) conducted an investigation of dissimilar metal joining combinations: FeTi, CuTi, FeCu, FeNi and CuNi. He observed the inuence of interaction time on microstructure and tensile properties of the friction welding of ve dissimilar metal combinations. Dey et al. (2009) carried out research concerning the friction welding of titanium to stainless steel. His studies conrmed the presence of secondary phases in the inter-mix zone near the interface. Seli et al. (2010) conducted both a mechanical evaluation and thermal modeling of the friction welding of mild steel and aluminum. They used

454

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

an explicit one-dimensional nite difference method to approximate the heating and cooling temperature distribution of the joint. zdemir (2003) studied the effect of the shape of the graphite in vacuum-free diffusion bonding of nodular cast iron with gray cast iron. From the results, he concluded that the shapes and surface areas of the graphite had an effect on the diffusion behaviors or the joining materials. zdemir (2005) also observed that the width of the fully plastic deformed zone (FPDZ) has an important effect on the strength of the friction-welded samples. Sunay et al. (2009) investigated the effects of casting and forging processes on joint properties in friction welded AISI 1050 and AISI 304 steels. Taban et al. (2010) joined aluminum and AISI 1018 steel applying friction welding. Arivazhagan et al. (2011) investigated AISI 304 stainless steel to AISI 4140 low alloy steel dissimilar joints by friction welding. Their friction processed joints were subjected to mechanical and microstructure investigations. The mechanical properties and microstructure of friction welded joints between AISI 4140 and AISI 1050 steel were also reported (Celik and Ersozlu, 2009). There were no cracks or blank spaces in the optical and SEM observations. The bonding of aluminum and ceramics was achieved in the friction welding process (Zimmerman et al., 2009). From the FEM calculation, it could be observed that during friction welding of ceramics with aluminum there were, in close proximity to the bond, uneven distributions of temperature, deformation and contact forces occur, which then cause inhomogeneity of the bond and inuence its strength. Reddy and Ramana (2012) have recently joined maraging steel to low alloy steel using nickel as an interlayer by friction welding. The study revealed that nickel can be employed as an effective barrier for the diffusion of elements at the interface. Friction welding joining is also suitable in the case of materials for which conventional welding is either very difcult or even impossible. For example, ductile iron can be successfully welded and also joined to other materials, such as steels with a high alloy-content. However, The American Welding Society (1989), and Lebedev and Chernenko (1992) from the Paton Electric Welding Institute, have concluded that the friction welding of ductile iron is not possible because graphite acts as a lubricant and prevents the generation of heat sufcient enough for joining. On the other hand, according to the Shinoda et al. (1996), from Japans Nagoya University, ductile iron can be joined by friction welding without any special treatment, such as preheating and/or post heating treatment. The tensile strength of friction-welded ferritic ductile iron specimens exceeds even 445 MPa, as reported by Shinoda et al. (1999). In this literature, studies concerning the friction welding of ductile iron with different materials can also be found. Richter and Palzkill (1985) conducted a study on the combination of constructional steel and spheroidal graphite cast iron. The authors concluded that in friction welding of steel with graphite-containing cast iron, the inuence of graphite in ductile iron on the welding process must also be taken into account, because this graphite builds up a lubrication layer which impedes the generation of an intensive frictional force and, consequently, the development of heat. A little later, Dette and Hirsch (1990) joined steel and spheroidal cast iron with friction welding. According to the authors, the main advantage to the applying of ductile iron was the weight saving of the part following the 10% lower gravity of ductile iron than standard steel. In order to attain the best welding quality, the carbide and alloying elements content should be kept to a minimum. Hareyama et al. (2007) studied friction-welded ductile cast iron pipes. In this study, the authors concluded that the tensile strength decreased with increasing layers of deformed spheroidal graphite, but high tensile strength specimens were broken on the base material. The results of the investigations concerning the microstructure and mechanical properties of friction-welded ductile cast iron were provided in a previous report by Winiczenko

(2001). Matsugi et al. (2004) investigated the inuence of joining conditions on joint properties by an impact-electric current discharge machine, and succeeded in obtaining a very strongly tensile spheroidal graphite cast ironstainless steel joint. Ogara et al. (2005) examined the relationship between tensile strength characteristics and the macrostructure of joints in friction-welded ductile cast iron. Ochi et al. (2007) studied the macrostructure and temperature distribution near the interface during the friction welding of FC250 grade cast iron. Their highest reported tensile strength in the solid joints and pipe joints were respectively 317 MPa and 381 MPa. Song et al. (2008) investigated the strength distribution at the interface of rotary-friction-welded aluminum to nodular cast iron. Nakamura et al. (2010) carried out research work about the inuence of the preheating temperature and welding speed of the microstructure of the joining zone obtained by friction stir welding (FSW) of ductile cast irons and stainless steels. The friction stir welding of ductile iron and low carbon steel was conducted also by Cheng et al. (2010). As can be seen in this literature, the main problem occurring in the friction welding of ductile iron is graphite having lubricating properties that reduce the efciency of the welding process. During the friction process, the graphite spheroids are deformed or fragmented, thus creating an unfavorable microstructure. High carbon content in ductile iron (more than 3.5%) constitutes a barrier for obtaining good quality joints. Very often this then leads to the formation of a hard and brittle martensitic microstructure in the heat affected zone (HAZ). In order to produce a good quality joint, many solutions are used. These include the optimization of welding parameters, the introduction of a low carbon steel interlayer, and changing the geometric shape of the joined parts; which are also heat treated before and/or after the friction welding process. Each of these cases provides many valuable insights to the friction welding process. Despite the previous reports about the poor weldability of ductile iron, Winiczenko and Kaczorowski (2012) succeeded in getting joints with a tensile strength of 700 MPa using interlayers. The evidence of the current studies illustrating that the microstructure of ductile iron should affect the quality of welded joints, due to the changes taking place when it is under the inuence of thermal effects, will shape the mechanical properties of joining materials (Metals Handbook, 9th ed.). This paper is a continuation of the experiment from which parts of the result were presented at Materials & Design 34 (2012) 444451 by its authors. Hence, the same equipment and similar materials were used. However, the current studies focused on the friction welding of ductile iron with a different metal matrix, using only stainless steel as an interlayer. The aim of this study is to dene the impact of the metal matrix of ductile iron on the microstructure and mechanical properties of the friction welded joints. Moreover, we would like to take a closer look into diffusion phenomena, accompanying the friction welding of ductile iron with stainless steel. As such, an EDS line, points, and the map spectrometry technique were used additionally.

2. Experimental procedure The chemical composition of test materials selected for the study is given in Table 1. Both a ferritic and bainitic matrix of ductile iron were produced, using different parameters of the heat treatment method. Microstructure bars, 20 mm in diameter and 100 mm in length, from a different casting, were cut as the specimens for friction welding. The surface for friction welding was prepared on the abrasive cut-off machine. The geometry of specimens used for

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462 Table 1 The chemical composition of materials (wt.%). Material Alloying elements (wt.%) C 60-45-12 AISI 321 3.503.90 0.08 max Si 2.253.0 0.8 max Mn 0.300.45 2.0 max P 0.0120.05 0.045 max S 0.0040.022 0.030 max Cr 0.030.2 1719 Ni 0.020.15 912 Mg 0.0410.105 Fe

455

Rest Rest

friction welding and details of the experiment were presented in Fig. 1. The friction welding of ductile iron specimens was conducted using AISI 321 stainless steel as an interlayer. The process of joining was carried out on the continuous drive friction machine of a ZT-14 type. Friction and forged pressures used in the experiment were in the range of 2045 kN. The spindle rotating speed was kept at a constant speed of 2360 rpm. Because of the presence of graphite, a relatively large friction time (FT) of 120270 s was applied. The tensile strength (TS) of the joints was determined using a conventional tensile test machine. Furthermore, the hardness across the interface was measured on a metallographic specimen. The Vickers tester, using the 0.05 N load, was applied for the hardness measurements. The microstructure of the joints was examined, using both light metallography as well as scanning electron microscopy (SEM). The latter one was carried out with a Jeol JSM-5400 scanning electron microscope. The surfaces of specimens were observed in the BEI COMPO mode, using back-scattered electrons (BSE). The studies were conducted on the interface using different variants including: EDSSEM map analysis and the EDSlinear analysis. In the rst case, the information was collected from the area whose size depended on the magnication. In this case, the EDS-analysis was performed for the distribution of such elements as C, Cr, Ni and Fe. The aim of the EDS-linear analysis was to determine the changes in the distribution of: C, Cr, Fe and Ni across the interface.

3. Results and discussion 3.1. Tensile test The tensile test was applied after machining the weld ashes, formed during the friction welding process. Ferritic and bainitic ductile iron microstructure samples were broken on the welding side of the ductile iron. The relationship between weld ashes of ductile iron joints and the friction time are given in Fig. 2a. As can be seen in Fig. 2a, ferritic ductile iron (FDI) diameter ashes were more plasticized than the bainitic ductile iron (BDI) specimens. The amount of weld ash joints increased with the increasing friction time for BDI joints, while weld ashes decreased with the increasing friction time for FDI and the parent matrix of ductile iron joints. Similar trends were achieved in the friction welding of dissimilar steels by Celik and Ersozlu (2009). In the case of dissimilar material joints by friction welding, the formation of the ash depends on the mechanical properties of the two parent materials. It was observed that the ashes were formed around the weld interface on the side of both ductile iron and stainless steels. The weld ash and the contact zone for FDI and BDI-stainless steel joints can easily be observed in Figs. 7b and 8b. The effect of the matrix of ductile iron on the tensile strength of welded joints has been presented in Fig. 2b. The samples were welded by constant welding parameters: friction force and the upsetting axial force. The friction time was changed from 150 to

Fig. 1. Shape and size of specimens (unit: mm) before friction welding: (a) friction welding of ductile iron to the interlayer as the rst welding cycle, (b) friction welding of ductile iron to ductile iron by the interlayer as a second welding cycle.

456

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

39,5

350

38,5

Tensile strength (MPa)

39
ferritic matrix parent matrix bainitic HF=UF=24 kN

300 250 200 150 100 50 0 100 200 300 400


ferritic matrix

Flash (mm)

38 37,5 37 36,5 36 0 100 200 300 400

parent matrix

bainitic matrix; HF=UF=24 kN

Friction time (s)

Friction time (s)

Fig. 2. The relationship between the friction time, ash diameter and tensile strength of friction welded ductile iron with various matrix microstructures: (a) ash diameter, (b) tensile strength.

300 s. Generally, the results of the tensile strength obtained for ductile iron-stainless steel specimens are not satisfactory. Smaller values for the average of tensile strength, equalling 195 MPa and 285 MPa, were obtained for the ferritic of the matrix ductile iron, whereas for the bainitic of the matrix ductile iron the values are much higher. The average of the tensile strength of the bainitic ductile iron reached 265 MPa. As can be observed from the diagram in Fig. 2b, the tensile strength of the joints increased with the bainitic matrix of the ductile iron, and the tensile strength decreased with a ferrite matrix for ductile iron and stainless steel joints. Similar results of the relationship between the matrix structure and tensile strength yet, instead, for ductile cast iron and pure Armco joints were achieved by Winiczenko and Kaczorowski (2012). The effect of friction time (FT) on the tensile strength of the various matrices of ductile iron is presented in Fig. 2b. As illustrated in the diagram, the tensile strength joints increased with the increasing friction time for BDI and stainless steel joints, but the tensile joints slightly decreased with the increasing friction time for FDI and stainless steel joints. 3.2. Hardness measurements Hardness measurements and the examination of metallurgical structures appearing at the given stage of the process provided valuable information on heat distribution and the microstructure produced by the heating and plastic deformation. The Vickers

microhardness distributions in specimens on both sides of the weld interface are shown in Fig 3. The measurements were carried out both in the central axis and along the line located 2.5 mm from the periphery (surface) for ferritic ductile ironstainless steel (Fig. 3a), and bainitic ductile ironstainless steel joints (Fig. 3b). As expected, the hardness reaches maximum close to the interface and decreases very rapidly in the stainless steel region (Fig. 3). Contrary to this, the hardness decrease in ferritic and bainitic ductile iron is much slower and exhibits some kind of plateau. As can be seen in the diagrams in Fig. 3a and b, this plateau extends to 5 or 8 mm from the interface, depending on the location of the hardness measurements, and reaches the value of 199 HV for BDI and 200 HV for FDI joints, which is typical of the parent material. The maximum hardness value obtained close to the interface reached 450 HV for BDI and 470 HV for the FDI joints. As expected, the hardness obtained in the weld zone of ferritic and bainitic ductile iron joints was much higher than the hardness of the parent material. The reason for this increase in the hardness at the weld zone is probably due to the creation of large amounts of carbides and other metallurgical products. The changes of micro hardness in the welding interface are directly associated with the microstructure formed and observed in the welding interface as a result of the degree of the heat input and plastic deformation (zdemir, 2003). The results of the microstructures of friction welded joined as function from the interface are described below.

a
500 450 400

b
500

ff

FDI

SS

FDI

450 400

BDI

SS

BDI

Microhardness HV 0,5

350 300 250 200

Microhardness HV 0,5

350 300 250 200

FDI 2.5mm

Bondline 1

Bondline 2

Bondline 1

100 50 0

100 50 0
35

15

95

55

75

55

35

15

55

05

95

45

65

.8

Bondline 2

150

FDI in axis

BDI 2.5mm BDI in axis

150

1. 55

7. 35

10

4.

2.

0.

3.

3.

4.

5.

5.

8.

9.

0.

Distance (mm)

Distance (mm)

Fig. 3. The hardness distributions of ductile ironstainless steel friction welded joints: (a) for ferritic ductile iron-FDI, and (b) for bainitic ductile iron-BDI.

10

.8 5

55

15

95

15

05

55

95

75

2.

0.

3.

5.

4.

3.

0.

1.

8.

4.

5.

9.

7.

35

45

65

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

457

Fig. 4. Examples of fracture surface morphology: (a) the brittle mode of a fracture throughout cleavage planes in ferrite surrounding graphite nodules, and (b) the mixed mode of fracture; shallow dimples visible in the central part of the micrograph.

Fig. 5. The deformed graphite in ductile iron. Magnication 3500.

3.3. Structure investigation The results of the SEM observations are given in Figs. 46. In the rst micrograph (Fig. 4), two examples of different morphology fracture surface were depicted. Fig. 4a shows the brittle one, while Fig. 4b illustrates the ductile mode of the fracture surface. In the rst photo, cleavage planes in a ferrite surrounding a graphite nodule are clearly visible. On the other hand, in Fig. 4b shallow dimples, which are characteristic for the ductile mode of a fracture, can be identied in the central area of the photo. The last observation could suggest the appearance of some amount of austenite in the ductile

iron microstructure. Fig. 5 shows highly distorted graphite nodules visible close to the joining plane. The results of the SEM observation of the ductile iron and stainless steel microstructure close to the frictionwelding interface are shown in Fig. 6. In the rst micrograph (Fig. 6a), carbide eutectic surrounding the graphite nodule is visible. In the next (Fig. 6b), carbides distributed at austenite grain boundaries in stainless steel can be detected. Semiquantitative EDS analysis showed the chromium carbides which are located at a distance of 2.5 mm from the ductile iron-stainless steel interface (Fig. 8d and h). The appearance of carbide eutectic in ductile iron can be explained only by the fact that the liquidsolid phase transformation had to proceed. Therefore, it can be proposed that the temperature of the surfaces of joining materials must (at least locally) exceed the melting point of ductile iron. The temperature in a double thermal effect was locally even over the melting point of ductile iron. This was conrmed by metallography that revealed the carbide eutectic enriched with Cr in ductile iron (Fig. 6a). The grain boundary carbides during diffusion welding of stainless steel with ductile cast iron were observed by (Kolukisa, 2007). The forming of carbides in stainless steel (Fig. 6b), which is almost carbon free, could be understood only in terms of carbon transportation from ductile iron to stainless steel through the interface. The presence of the carbide phase proves the diffusion of both C from ductile iron into stainless steel and alloying components from stainless steel into ductile iron. The diffusion of chromium and nickel atoms into ductile iron was also identied and reported. This process leads to the formation of Cr23 C6 type chromium carbides in ductile cast iron in the region very close to the ductile cast iron-stainless steel interface (Kaczorowski and Winiczenko,

Fig. 6. SEM micrographs showing the microstructure of joined materials: (a) carbide eutectic in ductile iron, (b) carbides in stainless steel located at the grain boundaries.

458

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

Fig. 7. EDS-analysis of Cr and Ni distribution across the interface in the axis and 7.5 mm from the center of ferritic ductile iron-stainless steel joint: (a) bondline 1 as rst cycle welding and bondline 2 as second cycle welding, (b) the weld ash macrograph with area of analysis, (c) the EDS spectrum obtained from FDI at the place marked as a green triangle in (b), and (d) from SS at the place marked as a yellow point in Fig 7b.

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

459

2013). According to many authors, this diffusion is the main bonding mechanism in the rotary friction welding process. The diffusion of chromium and nickel from stainless steel to ductile iron could explain the ductile mode of fracture shown in Fig. 4b. The formation of isolated areas of pearlite in an austenite matrix was observed in a previous report (Winiczenko, 2001). In other areas, approximately 2 mm from the interface, a grid of ferrite against a background of pearlite was revealed. It runs at the grain boundaries, which are easy paths for the diffusion of carbon atoms. 3.4. EDXSEM investigation Figs. 7 and 8 show the results of the EDS linear and point analysis of Cr and Ni distribution across the interface of FDI-SS (Fig. 7) and BDI-SS joints (Fig. 8) in the axis and at a distance of 7.5 mm from the axis of the joined samples for the rst and second welding cycle. The analysis of the records clearly indicates a diffusion of Cr and Ni through the interface from stainless steel to ductile iron. As a result of the diffusion, a gradual reduction of Cr and Ni in stainless steel while approaching the interface has been observed. A comparison between Cr and Ni distribution and that of carbon in particular places of the samples accommodates the earlier suggestion that during friction welding a diffusion of carbon from ductile iron into stainless steel occurs. Where joints subjected to a single welding cycle are concerned, its range equals 100 m in both the axis and at the periphery of the sample. In the double heat effect during the second cycle of welding, the depth of the carbon diffusion is greater and reaches up to 180 m for the area

at the periphery and 150 m in the centre of the sample. As we approach the interface, we observe a decrease in the average Cr concentration. In steel it starts at the distance of 110 and 120 m, respectively, for single and double thermal effects. The distribution of Cr has a mixed nature, whereas the size of localized peaks are larger for the joints that were subjected to double heat treatment. The range of the Cr diffusion to FDI is much smaller, and reaches only 50 m (Fig. 8). When taking into account the time of the friction welding, both the BDI and FDI was the same; the shorter range of Cr diffusion in the FDI could be explained by a lower temperature at the interface, which resulted from a lower amount of heat being produced during the friction process of FDI than BDI. The distribution of Ni is similar to that of Cr, although the width of the zone in which we observe its concentration changes in a function of distance on both sides of the interface is much higher. Moreover, on the side of the joint subjected only to friction welding, the decrease of Ni concentration starts at a distance of 90 m from the interface, while on the side of the joint subjected to annealing during the second cycle of welding the Ni concentration decrease starts at a depth of 120 m. As expected in the case of FDI, the gradual decrease of Ni concentration in stainless steel is observed at a much smaller distance (about 40 m). Fig. 8 shows the EDS spectrum, which makes possible the semiquantitative analysis at the points located close to the interface marked on the microstructure shown in Fig. 8c and d. Figs. 7c, d and 8eh illustrate the EDS spectrum taken from places 1, 2, 3 and 4 marked in the micrographs and showing the microstructure of BDI and SS close to the interface (Fig. 8c and d). The results of EDS

Fig. 8. EDS-analysis of Cr and Ni distribution across the interface in the axis and 7.5 mm from centre for the bainitic ductile iron-stainless steel joint: (a) bondline1 as rst cycle welding and bondline 2 as second cycle welding, (b) the weld ash macrograph with area of analysis, (c) the microstructure of BDI showing the rst and second places from where the EDS spectrum were taken, (d) the microstructure of stainless steel showing the third and fourth places from where the EDS spectrum was taken, (e) the EDS spectrum from BDI at the place marked as a green triangle in (b), (f) from SS at the place marked as a yellow point in (b), (g) the EDS spectrum from BDI near the interface at the rst place (left) and second (right) (see c and h), the EDS spectrum from SS near the interface at the third place (left) and fourth (right) (see d).

460

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

Fig. 9. The results of EDS analysis showing Ni, Cr, C and Fe distribution in the interface of a ferritic ductile iron-stainless steel joint.

analysis conrm that the precipitates visible on the SEM micrograph showing friction welded joints are chromium carbides. A more quantitative explanation of both atom and mass transport in friction welding has been published by the authors in the article: The Microstructure and Mass Transport during Friction Welding of

Ductile Cast Iron, Industrial Lubrication and Tribology, Vol.65, Issue 4, 2012. Figs. 9 and 10 show results of the EDS concerning the distribution of Ni, Cr, C and Fe at the interface of ferritic-stainless steel (Fig. 9) and bainitic stainless steel (Fig. 10) joints.

Fig. 10. EDS maps illustrating Ni, Cr, C and Fe distribution at the interface of a bainitic ductile iron-stainless steel joint.

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462

461

Using the results of these EDS studies as a basis, it can be concluded that, where ferritic ductile iron (FDI) is concerned, carbon occurs in the form of graphite clusters (nodules) only, while in bainitic ductile iron (BDI) carbon also forms as carbides and can then be treated as more uniform than FDI. During friction welding, the nodular graphite particles in the ductile iron are deformed to give ellipsoids, or are attened at least in certain places. These phenomena were observed by (Mitelea et al., 2010). The presence of graphite lubricated layers in the case of FDI, which may then be purely physical barriers that act on the principle of a discontinuity structure. The graphite bodies present at the interface in the form of clusters (Fig. 5) act as a lubricant and therefore produce insufcient heat from friction. Consequently, the properties of mechanical joints at the interface are reduced signicantly. The greater the amount of deformed graphite in the case of FDI, the thicker layer of graphite at the interface, and thus it is more difcult for both cross-diffusion and the possibility to create a joint.The depth of Cr atom penetration into ductile iron is substantially larger in ferritic ductile iron than in bainitic iron samples. Concentration of nickel in ductile iron with a bainitic matrix is larger than in ferritic matrix ductile iron. These observations are consistent with the results of the EDS linear microanalysis and conrm the results of microhardness changes at the interface. The results clearly showed that the friction welding process is inherent to the process of the transport of C, Cr and Ni atoms through the steel-ductile iron interface. As a result of this process there is an enrichment of cast iron with Ni atoms, which then forms a solid solution with iron and probably, at least locally, has a FCC lattice structure. At the same time, the diffusion of Cr leads to the formation of chromium carbides in stainless steel (Fig. 8d), located near the sample interface and in some cases found also as carbides in the carbide eutectic in ductile iron (Fig. 8c). It was also found that the intensity of the diffusion process was more pronounced during the friction welding of stainless steel with BDI rather than FDI. This is evidenced by a larger diffusion range, which is known to be directly related to the time and diffusion coefcient. Considering the identical time of friction welding of BDI and FDI, the difference in the depth of diffusion can only result from different diffusion coefcients. Although the metal matrix of friction welded ductile iron is different, it is highly unlikely that it could substantially affect the value of diffusion coefcients of the elements. It is, however, possible that the type of metal matrix affects the value of the friction coefcient. When considering that both ductile irons have the same content of C, it is easy to see that the FDI has a larger share of graphite than the BDI, in which at least 0.8% of carbon appears as a carbide phase. Moreover, it is known that graphite has excellent lubricating properties. If so, we can expect that the amount of heat generated during the welding of FDI may be substantially reduced. This in turn means less heat in the friction interface zone, thus a lower temperature and lower diffusion coefcient. A factor promoting a higher diffusivity of a bainitic matrix is the larger share of easy diffusion paths due to greater dispersion of a microstructure (a larger amount of grain and interphase boundaries). The high intensity of the diffusion process in friction welding of bainitic ductile iron (BDI) is caused by a much higher dislocation density, which results from the plastic deformation that takes place during friction welding. This increase of diffusivity is caused by pipe diffusion in the dislocation network of predeformed samples. Also, the penetration of diffusion is deeper in the plastic deformation material than in annealed material (as the rst welding cycle).Another explanation for the smaller range of diffusion of Cr and Ni from the SS to FDI, when compared to BDI, may be the larger shortening of FDI samples, due to the lower yield stress of FDI than BDI. If it is assumed that during friction welding both samples are heated to the same temperature, the plastic

deformation for the FDI joints will be greater than BDI when upset by the same force. A greater shortening of the specimens prevents metallic contact between the joining parts due to the material ow of FDI to ash, which would explain the smaller range of diffusion. However, both the shortening and diameter of ashes was less for FDI than BDI joints (Fig. 2a); and this above explanation cannot therefore be accurate. Moreover, the smaller shortening and smaller diameters of the weld ashes, where the friction welding of FDI is concerned, indicates a lower temperature at the interface than that obtained at the BDI joints. Another factor inuencing the diffusion process is the presence of graphite layers. This layer may act simply as a physical barrier in terms of a microstructure discontinuity. Graphite has excellent lubricating properties, resulting from easy sliding in hexagonal (0 0 0 1) planes. The easy sliding follows from the fact that these planes are bonded by very weak Van der Waals bonds (Metals Handbook, 8th ed., vol. 1). This greatly facilitates the distribution of graphite on the surfaces of welded elements. In addition to reducing the friction coefcient, a very thin graphite layer can substantially decrease the strength and even can make the bonding between joined parts impossible. If it is, therefore, assumed that there is a full join during friction welding then, after taking into consideration the share of graphite on the interface (around 10% of the surface in the case of FDI) and its almost zero tensile strength, it is easy to understand the negative effects of graphite on the tensile strength of the joints. This negative impact will be stronger the greater the share of graphite on the surface of the joined elements. 4. Conclusions The analysis of the microstructure, mechanical properties and distribution of elements across the interface of ductile iron and stainless steel allowed us to propose the following conclusions: Friction welding is accompanied by a transport of atoms in both directions across the ductile iron-stainless steel interface. This results in the enrichment of stainless steel with carbon, and ductile iron with chromium and nickel atoms. Stainless steel carbon enrichment results in the formation of chromium carbides that are distributed mostly at the grain boundaries. Iron enrichment in Cr and Ni resulted in the creation of an alloy ferrite. Cr was found also in a carbide eutectic. The range of Cr and Ni diffusion in iron generally does not exceed 50 m. The depth of the diffusion of carbon in the case of a joint subjected to a double thermal effect is 150 m and higher than for a sample subjected to one step friction welding. The intensity of diffusion processes during friction welding of bainitic ductile iron is larger than for ferritic ductile iron. Acknowledgment This work was supported by The State Committee for Scientic Research under Grant 7T08B05519. The authors wish also to thank professors Eugeniusz Ranatowski and Stanislaw Dymski from the Faculty of Mechanics of Bydgoszcz Technical University (Poland) for their valuable suggestions. References
American Welding Society, 1989. Specications and standards. In: Recommended Practice for Friction Welding. American Welding Society, Miami. Arivazhagan, N., Singh, S., Prakash, S., Reddy, G.M., 2011. Investigation on AISI 304 austenitic stainless steel to AISI 4140 low alloy steel dissimilar joints by gas tungsten arc electron beam and friction welding. Materials and Design 32, 30363050. Celik, S., Ersozlu, I., 2009. Investigation of the mechanical properties and microstructure of friction welded joints between AISI 4140 and AISI 1050 steels. Materials and Design 30, 970976.

462

R. Winiczenko, M. Kaczorowski / Journal of Materials Processing Technology 213 (2013) 453462 Reddy, G.M., Ramana, P.V., 2012. Role of Nickel as an interlayer in dissimilar metal friction welding of maraging steel to low alloy steel. Journal of Materials Processing Technology 212, 6677. Richter, H., Palzkill, A., 1985. Applicability of test result from miniature friction welded specimens to full-size specimens as demonstrated by the combination of constructional steel and spheroidal graphite cast iron. Welding and Cutting 37, 6065. Seli, H., Ismail A.I.Md. Rachman, E., Ahmad, Z.A., 2010. Mechanical evaluation and thermal modelling of friction welding of mild steel and aluminium. Journal of Materials Processing Technology 210, 12091216. Sahin, M., 2005. Joining with friction welding of high-speed steel and mediumcarbon steel. Journal of Materials Processing Technology 168, 202210. Sahin, M., 2009. Joining of stainless steel and copper materials with friction welding. Industrial Lubrication and Tribology 61 (6), 319324. Sahin Ahmet, Z., Yibas Bakir, S., Ahmed, M., Nickel, J., 1998. Analysis of the friction welding process in relation to the welding of copper and steel bars. Journal of Materials Processing Technology 82, 127136. Satyanarayana, V.V., Reddy, M.G., Mohandas, T., 2005. Dissimilar metal friction welding of austeniticferritic stainless steels. Journal of Materials Processing Technology 160, 128137. Shinoda, T., Endo, S., Tanada, K., 1996. Friction welding of cast iron and stainless steels. Welding Internship 10 (12), 929936. Shinoda, T., Endo, S., Kato, Y., 1999. Friction welding of cast iron and stainless steels. Welding Internship 13 (2), 8995. Song, Y., Liu, Y., Zhu, X., Yu, S., Zhang, Y., 2008. Strength distribution at interface of rotary-friction-welded aluminum to nodular cast iron. Transactions of Nonferrous Metals Society of China 18, 1418. Lancaster, J., 1987. Metallurgy of Welding. Allen and Unwin, London. Sunay, T.Y., Sahin, M., Altintas, S., 2009. The effects of casting and forging processes on joint properties in friction-welded AISI 1050 and AISI 304 steels. International Journal of Advanced Manufacturing Technology 44, 6879. Taban, E., Gould, J.E., Lippold, J.C., 2010. Dissimilar friction welding of 6061-T6 aluminum and AISI 1018 steel properties and microstructural characterization. Materials and Design 10, 23052311. Winiczenko, R., Kaczorowski, M., 2012. Friction welding of ductile cast iron using interlayers. Materials and Design 34, 444451. Winiczenko, R., 2001. The structure and properties of friction welded ductile iron. Warsaw University of Technology. Doctoral Dissertation. Warsaw, Poland. Zimmerman, J., Wosinski, W., Lindemann, Z.R., 2009. Thermo-mechanical and diffusion modelling in the process of ceramic-metal friction welding. Journal of Materials Processing Technology 209, 16531664.

Cheng, C.P., Lin, H.M., Lin, J.C., 2010. Friction stir welding of ductile iron and low carbon steel. Science and Technology of Welding & Joining 15 (8), 706711. Dette, M., Hirsch, J., 1990. Reibschweissen von Konstruieren aus Kugelgraphitguss mit Stahlteilen. Schweissen und Schneiden 42 (11), 188190. Dey, H.C., Ashfag, M., Bhaduri, A.K., Rao, K.P., 2009. Joining of titanium to 304L stainless steel by friction welding. Journal of Materials Processing Technology 209, 58625870. Kaczorowski, M., Winiczenko, R., 2013. The microstructure and mass transport during friction welding of ductile cast iron. Industrial Lubrication and Tribology 65 (4). http://www.emeraldinsight.com/journals.htm?articleid=17047124 Kolukisa, S., 2007. The effect of the welding temperature on the weldability in diffusion welding of martensitic (AISI 420) stainless steel with ductile (spheroidal graphite-nodular) cast iron. Journal of Materials Processing Technology 186, 3336. Lebedev, V.K., Chernenko, I.A., 1992. Welding and surface reviews. Friction Welding. E. O. Paton Electric Welding Institute. Matsugi, K., Konishi, M., Yanagisawa, O., Kiritani, M., 2004. Joining of spheroidal graphite cast iron to stainless steel by impact-electric current discharge joining. Journal of Materials Processing Technology 150, 300308. Metals Handbook, 8th ed., vol.1. Metals Handbook, 9th ed., vol. 6. ASM, Metal Park, Ohio, 1993. Meshram, S.D., Mohandas, T., Reddy, G.M., 2007. Friction welding of dissimilar pure metals. Journal of Materials Processing Technology 184, 330337. Mitelea, I., Craciunescu, C.M., Gugu, R., 2010. Interfacial behaviour of dissimilar friction welded nodular cast irons with low carbon steels. Materials Science Forum 638642, 37573762. Hareyama, T., Nitta, T., Kitagawa, M., Horie, H., 2007. Microstructure and mechanical properties of pipe shaped spheroidal graphite cast iron bonded by friction welding method. Journal of Japan Foundry Engineering Society 79 (3), 146150. Nakamura, M., Sawada, Y., Sato, Y.S., 2010. Metallographic study of lapped FSW between ductile cast iron and austenite type stainless steel. Materials Science Forum 638-642, 11971202. Ochi, H., Kawai, G., Morikawa, K., Yamamoto, Y., Suga, Y., 2007. Macrostructure and temperature distribution near the weld interface in friction welding of cast iron. Strength. Fracture and Complexity 5, 7988, 2009. Ogara, T., Kojoh, K., Nagayoshi, H., 2005. Relation between tensile characteristics and macrostructure of joint in friction welded ductile cast iron. Journal of Japan Foundry Engineering Society 77, 3943. zdemir, N., 2003. Effect of graphite shape in vacuum-free diffusion bonding of nodular cast iron with gray cast iron. Journal of Materials Processing Technology 141, 228233. zdemir, N., 2005. Investigation of the effect of the mechanical properties of friction welded joints between AISI 304L AISI steel as a function rotational speed. Materials Letters 59, 25042509.

Вам также может понравиться