You are on page 1of 9

Journal of Manufacturing ProcessesProcesses Journal of Manufacturing Vol. 6/No. 2 6/No. 2 Vol.

2004 2004

Two-Dimensional Finite Element Simulation of Material Flow in the Friction Stir Welding Process
Xiaomin Deng and Shaowen Xu, Dept. of Mechanical Engineering, University of South Carolina, Columbia, South Carolina, USA. E-mail: deng@engr.sc.edu

Abstract
Solid mechanics based finite element models and computational procedures have been developed by the authors to study and simulate the friction stir welding process. In this paper, two-dimensional simulation results of the material flow pattern and spatial velocity field around the rotating tool pin during welding, and the positions of material particles around the pin after welding, are presented. Material flow pattern predictions are found to compare favorably with experimental observations. Simulation results suggest that material particles in front of the tool pin tend to pass and get behind the rotating pin from the retreating side of the pin. Similarities between predicted velocity fields based on two different toolworkpiece interface models are described in detail, and implications of these findings (e.g., to fluid dynamics based models) are discussed.

Keywords: Friction Stir Welding, Material Flow, Finite Element Simulation, Solid Mechanics Model

Introduction
Friction stir welding (FSW) is a new solid-state joining process (Thomas et al. 1991) in which joining of material is achieved without melting and which has been found to be effective for joining hard-toweld metals, such as aluminum alloys (Dawes and Thomas 1996). As such, FSW has many advantages over traditional fusion welding and has been receiving increasing attention from the industry. Compared to other joining processes (e.g., gas metal arc welding), there exist relatively few studies of FSW in the literature. Most published papers on this subject (e.g., Mahoney et al. 1998; Murr, Liu, and McClure 1998; Reynolds and Duvall 1999) are devoted to experimental characterizations of material properties of friction stir welds. More recent work (Li, Murr, and McClure 1999; Colligan 1999; Reynolds 2000) has focused on observing material flow patterns in the FSW process.
Based on a presentation by the authors at the 29th North American Manufacturing Research Conference (NAMRC XXIX), 2001 (see Deng and Xu 2001).

Theoretical effort aimed at understanding the FSW process, such as computational modeling and simulation, has been limited mainly because of the complexity of this thermomechanical process (e.g., material flow, temperature rise, large plastic deformation, contact, and friction). Thermal models usually are focused on temperature prediction and neglect the material flow phenomenon. In particular, the papers by McClure et al. (1998) and Gould and Feng (1998) described analytical methods for computing the temperature field, and the study by Chao and Qi (1999) used the finite element method to obtain the temperature field solution. The paper by Russell and Shercliff (1999) presented an analytical model to determine the temperature field as well as microstructure evolution. Frigaard, Grong, and Midling (2001) employed the finite difference method and developed a three-dimensional heat flow model, which could also calculate the microstructure evolution and hardness distribution in friction stir welds. Several researchers have carried out fluid dynamics based simulations that include material flow effects. Smith et al. (1999) reported an effort to determine material properties (e.g., viscosity) for fluid dynamics simulations. Bendzsak, North, and Smith (1999) presented preliminary results from three-dimensional heat and material flow simulations in which viscous dissipation was the heat source (frictional heating between the rotating tool and the workpiece was not considered). Seidel and Reynolds (2001) described a two-dimensional simulation study based on fluid dynamics that predicted material flow patterns that compare well with experimental results. A three-dimensional rigid viscoplasticity model using computational fluid dynamics was carried out by Ulysse (2002) that provided a parametric evaluation of the effect of tool speeds on the temperature field.
125

Journal of Manufacturing Processes Vol. 6/No. 2 2004

Load Loading side Trading side

Tool shoulder

Tool pin

Figure 1 Schematic of Friction Stir Welding Process

modeling and simulation procedure for studying FSW, developed using a general-purpose commercial code, and presents results pertaining to the material flow pattern and spatial velocity field around the rotating tool pin during friction stir welding. As in any first attempt at gaining a progressive understanding of complex physical processes, the authors have made several simplifications and assumptions in this effort so that the problem becomes computationally tractable. These will be discussed in detail in subsequent sections.

Because of their difficulties, solid mechanics based models and simulations that include material flow effects appeared in the literature later than thermal or thermomechanical (non flow based) models and fluid dynamics models (which usually do not include consideration of elastic responses of the workpiece). In a short communication by Xu et al. (2001), a twodimensional modeling and simulation effort, based on the Arbitrary Lagrangian-Eulerian (ALE) finite element formulation and with material flow effects, was reported in which predicted marker positions compared well with experimental data. Details of the above effort and additional material flow results (e.g., velocity fields) were presented in Deng and Xu (2001), which serves as the basis of the current paper. The above effort was extended into the threedimensional case, and initial results were given in Xu and Deng (2002). Utilizing a hydro code, based on the finite difference method and a steady-state Eulerian formulation, Askari et al. (2001) studied three-dimensional material flow in friction stir welding. Dong et al. (2001) considered the plunging phase of the FSW process using the finite element method, in which material flow characteristics in the plunging phase were estimated using a finite difference based weld pool dynamics model. In addition to thermal, fluid dynamics, and solid mechanics models discussed above, there exists another type of analytic models (e.g., Nunes 2001) that utilizes insightful simplifications (e.g., kinematics assumptions) to derive expressions that describe in detail material flow characteristics of the FSW process. It seems that this model can yield spatial velocity field distributions comparable to those predicted by the current authors, as described below (also see Deng and Xu 2001). The current paper describes the details of a twodimensional solid mechanics based finite element
126

Problem Description
As shown schematically in Figure 1, this study deals with a butt weld that joins two identical plates made of Al 6061-T6 alloy. The tool pin is held between the two plates and rotates with an angular speed of and moves relative to the plates (or the plates move relative to the pin) with a constant velocity of v. A leading (or advancing) side and a trailing (or retreating) side can be defined. Joining of the plates is achieved through the combined action of friction heating generated at the shoulder-plate interface and the extrusion/forging effect created by the movement of the pin between two tightly held plates. As a first-order approximation, and to work within the bounds of current computational constraints, the FSW process is modeled as a two-dimensional problem. It is assumed that the plates to be joined are thick enough such that a state of plane strain can be achieved in the mid-plane. Because the tool is made of a material much stiffer than the plate material, the tool pin is taken to be rigid. Even though a fully coupled thermomechanical simulation procedure has been devised using a general-purpose commercial finite element code, in which deformation and material flow fields, as well as the temperature field, can be computed simultaneously, the limitation of our current PC-based computing power makes such a procedure impractical at this time. As such, the simulation procedure employed in this study is focused on determining the deformation and material flow fields. To compensate for the lack of a predicted temperature field, measured temperature values from an actual FSW test (McClure et al. 1998) are used to construct an approximate temperature field for the FSW process. This temperature field is then used as input for the solid mechanics

Journal of Manufacturing Processes Vol. 6/No. 2 2004

model for the same FSW process. As such, a problem geometry that accommodates the simulation of the preceding FSW test is used. Specifically, the radius of the pin is R = 3.25 mm, and the dimensions of the two plates are 100 mm in length (along the weld), 30 mm in width, and 6.4 mm in thickness. A tool rotation speed of 400 rpm and a plate translation speed of 2 mm/s are used in the FSW test. Temperature values in the test were measured at the mid-plane (3.18 mm below the surface) of the plates with thermocouples inserted in small holes drilled in the plates. The holes were aligned along a line normal to the welding direction. Temperature history values at various distances away from the weld line were then measured. Based on the measured values and steady-state condition, temperature variations along lines passing the measurement points and parallel to the weld line can be derived. These line variations are then fitted to functions along the lines, which are used to approximate the temperature field in the solid mechanics analysis. Figure 2 shows the fitted temperature history values.

Figure 2 Fitted Temperature History Values at Various Distances from Weld Line Table 1 Temperature-Dependent Material Properties for Al6061-T6 (Brown et al. 1993; Masubuchi 1980)

T (C) 25.00 100.00 148.89 204.44 260.00 315.56 371.11 426.67 482.22

E (GPa) 66.94 63.21 61.32 56.80 51.15 47.17 43.51 28.77 20.20

u (MPa) 278.12 260.68 251.24 221.01 152.26 73.87 36.84 21.58 10.49

0330 0.334 0.335 0336 0.338 0.360 0.400 0.410 0.420

Finite Element Model


In this study, the finite element simulation procedure for the FSW process is developed using various options in the general-purpose code ABAQUS. The two-dimensional geometry described above is divided into four-node quadrilateral elements. Reduced integration with hourglass control is used to avoid meshlocking problems associated with large incompressible plastic deformation. For a converged mesh, 30 rings of elements are used around the tool pin, with each ring containing 80 elements. The mesh consists of 24,000 elements and 24,460 nodes. The smallest elements are placed nearest to the tool pin boundary and have the dimension of 0.14 mm 0.25 mm. At the horizontal and vertical boundaries of the rectangular problem domain (see Figure 1), material particles move with a constant speed of v relative to the pin in the direction opposite to the translation movement of the pin. To accelerate the computation, the tool rotation and translation speeds are both increased 1000 times in the analysis, so that the ratio v/ R (which is 0.0147 in the test and represents the ratio of the plate moving speed relative to the pin and the tangential speed of the pin at its boundary due to pin rotation) stays the same as in the test. This acceleration is necessary
127

at this stage because simulation with the actual, slow speeds cannot be completed in a realistic period of time. To minimize the effect of change in absolute speed, the dependence of material behavior on the rate of deformation is eliminated in this analysis by treating the plate material (Al 6061-T6) as a rateindependent elastic-plastic material. However, the effect of temperature on yielding is considered explicitly in this analysis. True stress-strain curves and other properties for the material at different temperatures are shown in Figure 3 and Table 1 (where T is temperature, E is the Youngs modulus, u is the ultimate strength, and is the Poissons ratio). These data are obtained from the references by Brown, Mindlin, and Ho (1993) and Masubuchi (1980). Isotropic material behavior with isotropic hardening is considered. The constitutive relations are given by the von Mises yield criterion and the associated flow rule. Large deformation and material flow in the FSW process are handled with the adaptive remeshing and Arbitrary Lagrangian-

Journal of Manufacturing Processes Vol. 6/No. 2 2004

Figure 3 Temperature-Dependent True Stress-Strain Curves for Al 6061 T6 (Brown, Mindlin, and Ho 1993; Masubuchi 1980)

above which the frictional stress stays constant and is no longer equal to the product of the friction coefficient and the contact pressure. This modification is necessary in order to model the plastic shear flow behavior of the plate material when the applied shear stress is near the materials shear failure stress. In this study, max, is set to equal to u, where u is related to the materials ultimate strength u through the relation u = u/ 3 (based on the von Mises relationship between the yield stress in shear and yield stress in tension). Because u (thus u) has a range of values due to temperature dependence and because ABAQUS only allows a fixed value for max, an intermediate value will be chosen for max, as discussed later in more detail. In principle, because the frictional contact model treats contact and friction along the tool-workpiece interface explicitly, it is believed to be more realistic than the slipping interface model (including the special case of a sticking interface).

Eulerian (ALE) finite element formulation options in ABAQUS. Details of these options are included in ABAQUS manuals. Two different models are proposed for the pin-plate interface. In the slipping interface model, the pin boundary is treated as a slipping interface (it includes the sticking interface as a special case), such that plate material particles at the interface rotate with an angular velocity (say f) that is equal (the sticking interface case) or smaller than the rotating speed of the pin. This model is considered because it provides connections to possible fluid dynamics models (e.g., Seidel and Reynolds 2001), in which a boundary layer may develop along the pin boundary, or connections to solid mechanics models (e.g., Askari et al. 2001), in which a sticking interface is assumed so that the material at the interface rotates at the same speed as the tool. It seems that a so-defined slipping interface offers a simple way of simulating the effect of the boundary layer or a sticking interface. The disadvantage of this model is that the angular velocity f (or the ratio f / ) is not known in advanceit must be determined indirectly through comparisons with experimental results. In the frictional contact model, the interface between the pin and the plates may experience frictional contact described by a modified Coulomb friction law with a friction coefficient of , an option in ABAQUS. The Coulomb law is modified in the sense that there exists a maximum critical frictional stress, say max,
128

Results and Discussion


The two-dimensional simulation procedure has been used to model the FSW process with a range of process parameters while keeping the problem geometry, the converged finite element mesh, and the temperature-dependent material properties the same. The objectives of these simulations are: (a) to gain some understanding of the material flow and velocity distribution characteristics around the rotating tool pin during welding, (b) to see whether the plane strain model with a number of simplifying assumptions can capture the main features (at least qualitatively) of experimentally observed material particle positions around the pin, and (c) to gage the performance of the two proposed interface models. First, we look at the tangential velocity distribution at points along radial lines in several directions around the pin. Figure 4 shows the spatial variation of the tangential velocity based on the frictional contact interface model, and Figure 5 based on the slipping interface model, both for the case of v/ R = 0.0147. The scales in the figures are for the velocity only and not for the geometry. In Figure 4, = 0.3 is 7 the friction coefficient and max = 1.167(10 ) Pa is an 7 intermediate value between u = 1.012(10 ) Pa at 7 447C and u = 1.180(10 ) Pa at 432C, where 432C to 447C is the temperature range along the pin-plate interface, based on experimental temperature measurements, as discussed in a previous section. In Fig-

Journal of Manufacturing Processes Vol. 6/No. 2 2004

Figure 4 Tangential Velocity Distribution Along Radial Lines (Frictional Contact Interface Model)

Figure 6 Tangential Velocity Variation Along Pin-Plate Interface

Figure 5 Tangential Velocity Distribution Along Radial Lines (Slipping Interface Model)

Figure 7 Tangential Velocity Variation Along Circular Paths Around Pin-Plate Interface

ure 5, the reduced angular speed, f, at points surrounding the tool-pin interface is set to be 50% of the rotation speed of the tool pin, . The simulation results based on the two interface models show both similarities and differences. An obvious similarity is that both suggest the existence of a boundary layer near the interface where the spatial velocity is very high. The main difference also lies in the boundary layer. For example, while the speed at the interface is forced to be the same in the slipping interface model, it varies in the frictional contact interface model from the maximum at the bottom interface point (which is on the retreating or trailing side of the weld) to the minimum at the top interface point (which is on the advancing or leading side). To gain a better understanding of the similarities and differences between the two model results, addi129

tional figures are needed. Figure 6 shows the variation of the normalized tangential velocity, vt/v, along the pin-plate interface (denoted by the angular position ), which is at a distance of r = 3.25 mm from the pin center. The average speed along the interface from the frictional contact interface model is also shown. It is seen that this average speed is the same as the prescribed speed in the slipping interface model. Figure 7 shows the tangential velocity variations along circular paths of two different distances (r = 3.36 mm and r = 3.84 mm) away from the pin center, where stands for the angular position, as defined in Figure 6. It is clear that as r increases, the variations from the two interface models approach each other. In particular, at r = 3.84 mm (which is a distance of 0.59 mm from the pin-plate interface), the difference between the two simulation results is very small. This seems to suggest that the difference

Journal of Manufacturing Processes Vol. 6/No. 2 2004

Leading side

Rotating direction + Pin

Trailing side Plate moving direction

Figure 8 Post-Weld Marker Positions in a Friction Stir Welding Test (Reynolds 2000)

between the two models is not as large as they appear. Because of the lack of complete experimental data, quantitative verification of simulation results for velocity and material flow characteristics in FSW during and after welding is currently not possible. It appears that the only available test data in this regard are those obtained by our colleague at the University of South Carolina (Reynolds 2000). By using inserted material strips as marker materials (which have different compositions than the plate material) along the weld line, Reynolds was able to determine where material particles along a line perpendicular to the weld line are located after the tool pin has passed through. For example, Figure 8 shows the post-weld marker positions in the mid-plane of a weld. In this FSW test, the tool pin has a radius of 5 mm and rotates at the speed of 215 rpm, and the plates move relative to the pin at a speed of 2.35 mm/s. The plate material is aluminum alloy 2195-T8. It is noted that there are differences (e.g., in plate material and speeds) between the test that supplied this study with temperature measurements (McClure et al. 1998) and the test shown in Figure 8. Because so far no one published FSW test with the same material and geometry properties has both the temperature and the marker position measurements, the current authors are not able to conduct a direct and quantitative validation of the present numerical simulation procedure. However, if we accept that the marker positions shown in Figure 8 represent a common material flow pattern in FSW, a qualitative comparison of the simulation results with the test data

can then be made. As such, the main purpose of the comparison is to see if the solid mechanics simulation procedure can capture the main features of material flow pattern in FSW. To this end, simulation marker positions shown in Figures 9 and 10 are examined. In these figures, the predicted marker positions are indicated by the crossed squares, each representing the position of a material particle after the rotating pin has past. These particles were originally aligned along a straight narrow band perpendicular to the weld line. The contour lines around the pin represent the spatial flow velocity distribution around the pin. A good qualitative agreement between the measured and simulated marker positions is observed. The ability of the plane strain simulation procedure to capture an experimentally observed material flow pattern in FSW was first reported in a brief technical note by the current authors and their collaborators (Xu et al. 2001) and represents an important step toward the development of a three-dimensional simulation procedure for the FSW process. One advantage of the developed computer simulation tool is that it allows us to understand and visualize the effect of process parameters (such as tool translation and rotation speeds) on the FSW process. The same ability is not always readily achievable or economical with experimental techniques. For example, Figures 11 and 12 show the positions of the markers in Figures 9 and 10 when the tool rotation speed is reduced 10 times (note that the ratio f / in Figure 11 is 1 instead of 0.5 as in Figure 9, which is used to further demonstrate a similarity in the flow patterns produced by the two interface models). It is seen that the marker band above the pin is now bent toward the plate moving direction. This behavior is intuitively correct because now, with a reduced rotation speed, the marker materials moving with the plate experience less counter motion exerted on them by the rotation of the pin (note that the tangential velocity due to rotation is opposite to the plate velocity relative to the pin). In both Figure 11 and Figure 12, the contours around the pin provide a view of a steady-state spatial velocity distribution during the FSW process. The similarity between the velocity fields produced by the two interface models explains the similarity in the post-weld marker positions. In particular, a sense of how the tangential velocity varies in space along

130

Journal of Manufacturing Processes Vol. 6/No. 2 2004

Figure 9 Predicted Post-Weld Marker Positions (Slipping Interface Model)

Figure 11 Predicted Post-Weld Marker Positions When Rotation Speed is Reduced 10 Times (Slipping Interface Model)

Figure 10 Predicted Post-Weld Marker Positions (Frictional Contact Interface Model)

Figure 12 Predicted Post-Weld Marker Positions When Rotation Speed is Reduced 10 Times (Frictional Contact Interface Model)

the radial direction can be gained from Figures 13 and 14. It is seen that, with reference to Figures 4 and 5, the boundary layer around the pin is now thicker, probably because of a more diminished dominance of the rotation effect due to a reduced rotation speed, but the similarity between the two flow fields is stronger in this case. A more quantitative comparison between the flow fields predicted by the two interface models is given in Figures 15 and 16, in which the tangential velocity variation along circular paths around the tool pin is plotted as a function of the angular position (see definition in Figure 6). The distance of the paths from the center of the pin is indicated in the figures. Several observations can be made from the velocity fields shown in Figures 15 and 16 as well as in figures shown earlier (Figures 4, 5, 7, 13, and 14). First, although in the slipping interface model material particles along the interface are made to rotate with the pin with the same speed, away from the interface

the spatial velocity distribution tends to approach that produced by the frictional contact interface model. This suggests a convergence of the two different modeling approaches. On one hand, a frictional contact interface model is a natural choice for solid mechanics problems involving contact and friction, even though a prescribed interface velocity boundary is much simpler. On the other hand, fluid dynamics models cannot treat contact and friction but they can readily handle prescribed velocity boundary conditions along the pin-plate interface, whether the prescribed velocity is the same as or smaller than the velocity due to the actual pin rotation. This observation thus provides an important basis for the fluid dynamics based modeling effort for the FSW process. Second, a phenomenon that is not intuitive and not immediately apparent from the figures is the manner in which material particles flow past the rotating pin as the plate moves relative to the pin. The authors have used the visualization capability of the simulation tool and produced movies showing

131

Journal of Manufacturing Processes Vol. 6/No. 2 2004

Figure 13 Tangential Velocity Distribution Along Radial Lines When Rotation Speed is Reduced 10 Times (Frictional Contact Interface Model)

Figure 15 Tangential Velocity Variation Along Circular Paths Around Pin-Plate Interface (Slipping Interface Model)

Figure 14 Tangential Velocity Distribution Along Radial Lines When Rotation Speed is Reduced 10 Times (Slipping Interface Model)

Figure 16 Tangential Velocity Variation Along Circular Paths Around Pin-Plate Interface (Frictional Contact Interface Model)

clearly how marker materials in front of the pin move toward the pin and flow around the pin. What we have observed from the movies is that, when marker materials directly in front of the rotating pin pass the pin, they do not get around the pin from both sides of the pin. Rather, they tend to rotate with the pin when they approach the pin-plate interface and will always get behind the pin from the trailing (retreating) side of the pin (which is the side in the lower portion of Figures 4, 5, 13, and 14). Only material particles near the very edge of the leading (advancing) side of the pin (corresponding to = 90) will move past the pin without first rotating with the pin. This material flow pattern is now also clear from Figures 4, 5, 7, 13, 14, 15, and 16. These figures all indicate that the tangential velocity field around the pin
132

is geared toward the rotating direction of the pin, except very near = 90, where the tangential velocity is reduced to zero or becomes slightly negative. An additional observation from the movies is that material particles in the leading side of the weld line before welding will tend to stay together in the leading side after welding, and those in the trailing side will tend to stay together in the trailing side, perhaps with some overlapping across the weld line. Measured marker positions (see Reynolds 2000) have confirmed this simulation result.

Concluding Remarks
A plane-strain finite element procedure has been developed to simulate the friction stir welding (FSW) pro-

Journal of Manufacturing Processes Vol. 6/No. 2 2004

cess focusing on velocity field and material flow characteristics. Two models have been proposed for the pinplate interface. Simulation predictions of post-weld marker positions based on both models compare well with experimental measurements. In conclusion, the authors would like to stress that, due to computational constraints and the lack of experimental data, several simplifications have been made in carrying out this study to make the problem tractable. As such, the comparison in this paper with test results is meant to demonstrate the capability of the simulation procedure to capture the essential features of the FSW process, and not to provide a quantitative prediction at this stage. However, work is under way by the authors to eliminate the simplifications. As experimental data and computational resources become available in due time, more accurate and realistic simulations of the FSW process will become available. Acknowledgments This work was sponsored by the National Science Foundation (grant no. DMI-9978611). Discussions with Professor A.P. Reynolds, Professor Y.J. Chao, Dr. W. Tang, and Mr. T.U. Seidel are gratefully acknowledged. References
Askari, A.; Silling, S.; London, B.; and Mahoney, M. (2001). Modeling and analysis of friction stir welding processes. Friction Stir Welding Processing, K.V. Jata, M.W. Mahoney, R.S. Mishra, S.L. Semiatin, and D.P. Field, eds. Warrendale, PA: The Minerals, Metals & Materials Society (www.tms.org), pp43-54. Bendzsak, G.J.; North, T.H.; and Smith, C.B. (1999). An experimentally validated 3D model for friction stir welding. Proc. of 1st Intl Symp. on Friction Stir Welding, Thousand Oaks, CA, June 1999. Brown, W.F., Jr.; Mindlin, H.; and Ho, C.Y. (1993). Aerospace Structural Metals Handbook (v3). CINDAS/Purdue University. Chao, Y.J. and Qi, X. (1999). Thermal and thermo-mechanical modeling of friction stir welding process. Journal of Materials Processing and Mfg. Science (v7, n2), pp215-233. Colligan, K. (1999). Material flow behavior during friction stir welding of aluminum. Welding Journal (July 1999), pp229-237. Dawes, C.J. and Thomas, W.M. (1996). Friction stir process welds aluminum alloys. Welding Journal (v75, n3), pp41-45. Deng, X. and Xu, S. (2001). Solid mechanics simulation of friction stir welding process. Transactions of NAMRI/SME, Vol. XXIX. Dearborn, MI: Society of Manufacturing Engineers, pp631-638. Dong, P.; Liu, F.; Hong, J.K.; and Cao, Z. (2001). Coupled thermomechanical analysis of friction stir welding process using simplified models. Science and Technology of Welding and Joining (v6, n5), pp281-287. Frigaard, O.; Grong, O.; and Midling, O.T. (2001). A process model for friction stir welding of age hardening aluminum alloys. Metallurgical and Materials Trans. A: Physical Metallurgy and Materials Science (v32, n5), pp1189-1200. Gould, J.E. and Feng, Z. (1998). Heat flow model for friction stir welding of aluminum alloys. Journal of Materials Processing and Mfg. Science (v7), pp185-194.

Li, Y.; Murr, L.E.; and McClure, J.C. (1999). Solid-state flow visualization in the friction-stir welding of 2024 Al to 6061 Al. Scripta Materialia (v40), pp1041-1046. Mahoney, M.W.; Rhodes, C.G.; Flintoff, J.G.; Spurling, R.A.; and Bingel, W.H. (1998). Metallurgical and Materials Trans. A (v29A), pp1955-1964. Masubuchi, K. (1980). Analysis of Welded Structures. Oxford, UK: Pergamon Press. McClure, J.C.; Feng, Z.; Tang, W.; Gould, J.E.; Murr, L.E.; and Guo, X. (1998). A thermal model of friction stir welding. Proc. of 5th Intl Conf. on Trends in Welding Research, Pine Mountain, GA, June 1998. Murr, L.E.; Liu, G.; and McClure, J.C. (1998). A TEM study of precipitation and related microstructures in friction-stir-welded 6061 aluminium. Journal of Materials Science (v33), pp1243-1251. Nunes, A.C., Jr. (2001). Wiping metal transfer in friction stir welding. Aluminum 2001: Proc. of 2001 TMS Annual Meeting, Automotive Alloys and Joining Aluminum Symposia, G. Kaufmann, J. Green, and S. Das, eds. Warrendale, PA: The Minerals, Metals & Materials Society (www.tms.org), pp235-248. Reynolds, A.P. (2000). Visualisation of material flow in autogeneous friction stir welds. Science and Technology of Welding and Joining (v5, n2), pp120-124. Reynolds, A.P. and Duvall, F. (1999). Welding Journal Research Supplement (Oct. 1999), pp355-360. Russell, M.J. and Shercliff, H.R. (1999). Analytical modeling of microstructure development in friction stir welding. Proc. of 1st Intl Symp. on Friction Stir Welding, Thousand Oaks, CA, June 1999. Seidel, T.U. and Reynolds, A.P. (2001). A two-dimensional fluid mechanics based friction stir welding process model. 12th Annual Advanced Aerospace Materials & Processes Conf. and Exposition (Aeromat 2001), Long Beach, CA, June 2001, pp11-14. Smith, C.B.; Bendzsak, G.B.; North, T.H.; Hinrichs, J.F.; Noruk, J.S.; and Heideman, R.J. (1999). Heat and material flow modeling of the friction stir welding process. Proc. of 9th Intl Conf. in Computer Technology in Welding, Detroit, MI, Sept. 1999. Thomas, W.M., et al. (1991). Friction stir butt welding. U.S. Patent No. 5,460,317. Ulysse, P. (2002). Three-dimensional modeling of the friction stir-welding process. Intl Journal of Machine Tools & Manufacture (v42), pp1549-1557. Xu, S.; Deng, X.; Reynolds, A.P.; and Seidel, T.U. (2001). Finite element simulation of material flow in friction stir welding. Science and Technology of Welding and Joining (v6, n3), pp191-193. Xu, S. and Deng, X. (2002). A three-dimensional model for the frictionstir welding process. Paper #2108. Proc. of 21st Southeastern Conf. on Theoretical and Applied Mechanics (SECTAM XXI), Orlando, FL, May 2002, pp699-704.

Authors Biographies
X. Deng is a professor of mechanical engineering at the University of South Carolina. He received his BS degree in 1982 from the Beijing University of Aeronautics and Astronautics (China) and his MS (1985) and PhD (1990) from the California Institute of Technology. His current research interests include modeling and simulation of manufacturing (machining and welding) processes, 3-D mixed-mode fracture criterion and simulation code development, and nanomechanical analysis using molecular dynamics simulations. S. Xu is a post-doctoral research associate in the Dept. of Mechanical Engineering at the University of South Carolina. He received his BS (1985) and MS (1992) degrees from the Huazhong University of Science and Technology (China) and his PhD (2003) from the University of South Carolina. His research interests include finite element modeling and simulation of engineering systems and manufacturing processes, nanomechanical simulations, and experimental material characterization of welded joints (e.g., friction stir welds).

133

You might also like