You are on page 1of 7

Scripta Materialia, Vol. 37, No. 3, pp.

355-361, 1997
Elsevier Science Ltd
Copyright 6 1997 Acta Metahrgica Inc.
printed in the USA. All rights resewed
1359-6462/97 $17.00 + .OO
PII S1359-6462(97)00093-6

MICROSTRUCTURAL ASPECTS OF THE


FRICTION-STIR WELDING OF 6061-T6 ALUMINUM
G. Liu, L.E. Murr, C-S. Niou, J.C. McClure, and F.R. Vega
Department of Metallurgical and Materials Engineering
The University of Texas at El Paso, El Paso, TX 79968
(Received November 2 1,1996)
(Accepted February 2 1, 1997)

Introduction

Friction-stir welding is a derivative of conventional friction welding (1). In the friction-stir welding pro-
cess, a so-called welding-head pin rotating at speeds usually in excess of a few hundred rpm, travels down
the length of contacting metal plates, creating a highly plastically deformed zone through the associated
force and frictional heating. This plastic deformation zone is essentially “stirred” into a solid-phase weld
on the trailing side of the welding head pin. This process can also be simulated in a solid metal plate by
simply advancing the rotating head pin into the work piece; with a residual weld zone forming on its trail-
ing side. Figure, 1 illustrates the friction-stir welding process.
As shown schematically in Fig. 1, a rotating steel pin (HP) advances into the workpiece, creating a
highly deformed, plastic zone which flows around to its trailing side. No melt occurs, and the weld forms
by solid-state plastic flow at elevated temperature. There is little or no porosity or other, related fusion
weld-type defects associated with the weld zone if the rotational speed (R) and traverse speed (T) are
optimized. The simplicity and applicability of the concept to even dissimilar materials pose a wide range
of industrial/mamufacturing prospects (2-4).
In this study, we have used light metallography (microscopy) (LM) and transmission electron micro-
scopy (TEM) to characterize the microstructures in the friction stir weld zone (and including the transition
region), and calmpare them with the original 6061-T6 aluminum alloy work-piece microstructures. We
have also measured the associated microhardness profile extending from the workpiece and through the
weld zone. It is therefore the intent of this paper to present a very brief but comprehensive microstructural
overview of this process and illustrate corresponding hardness profiles associated with these microstruc-
tures. We are not aware of any previous, TEM microstructural comparisons of a friction stir weld.

ExDerimentaI Details

6061-T6 aluminum alloy plate (nominally 0.63 cm thick) was used in a series of friction-stir welding
experiments to ‘bereported herein. A series of butt welds and simulated welds in solid plate sections were
conducted, as illustrated schematically in Fig. 1, at rotational speeds (R) ranging from 300 to 1000 rpm,
and translational (traverse) velocities of 0.15 to 0.25 cm/s. The hardened carbon steel welding head pin
(HP in Fig. 1) was 0.63 cm in diameter and its length was .58 cm. The shoulder/chuck diameter was

355
356 FRICTION-STIR WELDING Vol. 37, No. 3

Figure 1. Schematic arrangement illustrating friction stir welding in thin plate. A rotating (R), hard, head-pin (or nib
shown at HP) having dimensions approximately equal to the plate thickness advances into the workpiece at some selected
traverse speed (T) to create a solid-phase weld in its trailing side.

1.25 cm. Welded cross-sections were ground, polished, and etched with Keller’s reagent (150 mL water,
3 mL nitric acid, 6 mL hydrochloric acid, 6 mL hydrofluoric acid; at O°C) for optical metallography. Thin
specimens were sliced from these cross-sections at various locations within and outside the weld zone,
ground to 0.2 urn thickness, mechanically dimpled, and 3 mm discs punched for transmission electron
microscopy (TEM). A Tenupol-3 dual jet electropolisher was used to produce electron transparent thin
sections in these punched discs using a 20% nitric acid solution in methanol at -20°C. Transmission
electron microscopy was performed in a Hitachi H-8000 instrument operated at 200 kV accelerating
voltage, and employing a goniometer-tilt stage. Energy-dispersive X-ray spectrometry was also employed
for precipitate analysis. Instrumental (digital) Vickers microhardness measurements were also made
throughout the weld zone and into the initial aluminum alloy plate using a 100 gf load.

Results and Discussion

Figure 2 illustrates a low magnification overview of a typical friction stir weld zone in 606 1-T6 aluminum
alloy plate along with corresponding, Vickers hardness profiles. Figure 2(c) also shows a magnified view
of the weld zone transition at A in Fig. 2(a). Several features are prominent in Fig. 2. The weld zone wi-
dens near the upper surface in contact with the rotating head-pin fixture. The frictional heating is greater
near the top surface where maximum temperatures have been measured near 250°C. The vertical hardness
traverse in Fig. 2(a) shows this effect as slight softening relative to the weld zone base. That is, hardness
values near the top of the weld in Fig. 2(a) are between 50 and 60 VIIN in contrast to 52 to 70 VI-IN near
the bottom of the weld as shown in Fig. 2(b). There is also considerable softening throughout the weld
zone compared to the workpiece/plate hardness (Fig. 2(b) even though the average grain size is reduced
by a factor of about 10 (from about 100 pm in the workpiece to 10 pm in the weld zone) (Fig. 2(c)). This
phenomenon is a characteristic of dynamic recrystallization which appears to provide a mechanism for
solid-state, plastic flow facilitating the actual weld and weld-zone development.
Figure 3 compares the weld zone and workpiece grain structures and microstructures by LM and TEM.
The TEM bright-field images in Fig. 3(b) and (d) illustrate that although there has been a reduction in the
grain size by dynamic recrystallization (as well as a change from slightly elongated grams in the plane of
the workpiece to more equiaxed grains in the weld), there is a reduction in dislocation density. In addition,
there is a complex variation in the precipitation especially associated with the transition zone connecting
the work piece or undisturbed original plate with the residual weld zone. Figure 4 attempts to provide an
overview of the precipitation phenomena. Figure 4(a) and (b) compare the weld zone center precipitation
with the workpiece far removed, respectively. These homogeneously distributed precipitates (with average
stoichiometries of roughly Al,MgSi; x = 8 to 10) are generally smaller in the workpiece zone than in the
weld zone. However, there are far fewer large precipitates in the weld zone (compare Fig. 3(a) and (c)).
Vol. 37, No. 3 FRICTION-STIR WELDING 357

Figure 2. Friction-stir weld zone cross-section produced at a head-pin rotational speed of 400 rpm and a traverse velocity of 0.21
cm/s (see Fig. 1). (a) Low-magnification section (thickness) view. The arrow marked (R) indicates the direction of head-
pin rotation. (b) Vickers microhardness traverse through the section in (a) along horizontal reference line shown dotted (in (a));
and vertical profile indicated in (a). (c) Magnified view of weld zone and weld transition into the alloy plate in location shown
at A in (a).
358 FRICTION-STIR WELDING Vol. 37, No. 3

Figure 3. Comparison of LM and TEM images of microstructures in the 6061-T6 aluminum alloy workpiece and in the friction-stir
weld zone. (a) LM of workpiece grain structure. (b) TEM of workpiece microstructure showing heavy dislocation structure. The
selected-area electron difhaction pattern insert shows a [112] zone axis with g = [I lr]. (c) LM of weld zone grain structure at same
magnification shown in (a). (d) TEM of weld zone microstructure showing precipitates and dislocation structure at same
magnification shown in (b).
Vol. 37, No. 3 FRICTION-STIR WELDING 359

Large precipitams contained a considerable amount of iron, and precipitates in grain boundaries such as
those shown in Fig. 3(d) had a composition of AlSi. More crystallographic or lath-type precipitates occur-
ring in the upper transition zone regions, as shown typically in Fig. 4(c) and (d) were similar in compo-
sition to the homogeneously distributed precipitates, except they usually contained copper in roughly the
same proportion as magnesium and silicon (Al,MgSiCu). Near the bottom of the weld and just outside the
transition region (marked (4e) in Fig. 2(a)) coherent, GP-type precipitates on { 100) cube planes were
observed as shown in Fig. 4(e). Thicker, noncoherent precipitates seemed to evolve in other regions espec-
ially near the center of the workpiece in the transition zone as shown in Fig. 4(f), which also illustrates
dislocations and B high density of dislocation loops associated with these precipitates.
In classical precipitation evolution in systems like Cu-Al alloys, the range of precipitate phases from
GP- 1 zones to 13,for example, characterize an aging temperature range of more than 50°C. However in
this system, the precipitation is not well documented, and corresponding aging temperatures are unknown.
Nonetheless, the: complex range of precipitation shown in Fig. 4 attests to the fact that a temperature
gradient exists from the top to the bottom of the weld as well as in the transition region connecting the
weld zone to the workpiece. It will require considerably more detailed studies involving temperature
variations in friction-stir welded samples of 6061-T6, and especially variations of microstructures
(particularly precipitation microstructures) with rotation and translation speeds to more fully understand
their relationships and evolutionary features.
While, as noted on perusing Figs. 2 and 3 in retrospect, dynamic recrystallization seems to play a signi-
ficant role in the friction-stir welding process as studied in this work, the “recrystallization” may not be
classical. In the more classical context, recrystallization proceeds by nucleation and growth of new grains
surrounded by high-angle boundaries. This process is often referred to as dynamic discontinuous recry-
stallization and can include “cycles” of deformation building to a critical value of stored energy, creating
nuclei which grow, are deformed, and the process continues (3,6). These kinds of processes are attributed
to accommodating the extreme, superplastic deformation required for jet development in shaped charges,
and the flow of material during cratering and high velocity penetration (7- 10).
It is interesting to note that in the friction stir weld development illustrated and compared in Figs. 2
and 3, the grain s,ize within the weld zone is not extremely small, and as a consequence simple concepts
involving grain blotmdary sliding to accommodate the superplastic deformation associated with the friction
stir weld seem inadequate. However, considerable grain growth might occur after recrystallization. Plastic
strain incompatibility is not only caused by grain boundary sliding, but also by slip and cross-slip. When
second-phase particles occur as in this experimental study (Figs. 3 and 4), intergranular accommodation
slip must also occur to satisfy the surface traction continuity and displacement at these interphase boun-
daries during the deformation.
In friction stir weld development, slip deformation appears to dominate, and recrystallization seems
to evolve by dynamic continuous recrystallization (6,11- 13) where sub-grain boundaries form and incur
a continuous increase in misorient&ion. The process is therefore characterized by the introduction of ac-
commodation dislocations within subgrains, as shown in Fig. 3(b) and their absorption into the boundaries
by “annealing”, promoted by friction-induced heating in the weld zone. Some insight into the mechanism
of dynamic continuous recrystallization during superplastic deformation is provided in a recent paper by
Tsuzaki, et al. (14).

Conclusions

The friction-stir weld zone in 606 1-T6 aluminum is characterized by what appears to be a dynamic con-
tinuous recrystallization microstructure. Second-phase particles in the workpiece are essentially “stirred”
into the weld zone where the residual hardness varies from 55 WIN near the top of the weld to 65 WIN
near the bottom; in contrast to a workpiece hardness which varies between about 85 and 100 WIN. The
360 FRICTION-STIR WELDING Vol. 37, No. 3

Figure 4. Comparison of precipitation phenomena associated with the friction-stir weld. (a) Homogeneous precipitation in
the friction-stir weld zone at (4a) in Fig. 2(a). (b) Homogeneous precipitation in the workpiece far removed from the weld zone
(a) and (b) observed by tilting for no diilocation contrast in the TEM. Very large particles observed in Fig. 3(a) are shown by arrow
in (b). (c) Lath-like precipitates in (100) cube planes in upper letI transition region of weld zone (marked (4~) in Fig. 2(a)). (d)
Lath-lie precipitates in upper right portion of weld zone marked (4d) in Fig. 2(a). (e) Coherent - ( 100) precipitate zones near weld
zone bottom marked (4e) in Fig. 2(a). (f) Thin, noncoherent precipitates on crystallographic planes near the workpiece
center transition zone (marked (49 in Fig. 2(a)).
Vol. 37, No. 3 FRICTION-STIR WELDING 361

weld zone grain size averaged 10 pm in contrast to 100 pm for the workpiece. However, complex precipi-
tation phenomena associated with the transition zone extending between the workpiece and the actual weld
zone illustrate the need to study this process in considerably more detail.

This research was supported in part by a NASA-Marshall Space Flight Center Grant NAG 8- 1056, and
in part by a Mr. and Mrs. Macintosh Murchison Chair Endowment (L.E.M.) at the University of Texas
at El Paso.

References

1. H.B. Cary, Modem Welding Technology, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1979, p. 223.
2. W.M. Thomas, et al. “Friction Stir Butt Welding,” International Patent Application No. PCT/GB92/02203 and GB Patent
Application No. 9125978.8, Dec. 6, 1991.
3. O.T. Midling, “Material Flow Behavior and Microstmctural Integrity of Friction Stir Butt Weldments,” Proc. 4th Int. Conf.
on Aluminum Alloys (ICAA4), Atlanta, GA, USA, 1l-16 September, 1994.
4. C.J. Dawes, ‘An Intmduction to Friction Stir Welding and Its Development”, Welding & Me&l Fab., January, 1995, p. 12.
5. H.J. McQueen, J.P. Bailon, J.I. Dickson, and J.J. Jones (editors), ICSMA-7, Pergamon Press, Oxford, 1986.
6. C.H. Hamiltnn and N.E. Paton (editors), SuperpZasticityandSupetplastic Forming, Min. Metals & Mater. Sot., Philadelphia,
PA, 1988.
7. L.E. Murr, H.K. Shih, C-S. Niou, Mater. Characterization, 33,65 (1994).
8. J.M. Rivas, S.A. Quinones, and L.E. Murr, Scripta Metall. et Mater., 33(l), 101 (1995).
9. S.A. Quinones, J.M. Rivas, E.P. Garcia, L.E. Murr, F. H&r., and R.P. Bernhard, Chap. 36 in h4etullurgical and Materials
Applications of Shock Wave and High-Strain-Rate Phenomena, L.E. Murr, K.P. Staudhammer, and M.A. Meyers (Eds.),
Elsevier Science, B.V., Amsterdam, 1995, p. 313.
10. L.E. Murr, C-S. Niou, E.P. Garcia, E. Ferreyra T., J.M. Rivas, and J.C. Sanchez, “Comparison of Jetting-Related
Micmstmctures Associated with Hypervelocity Impact Crater Formation in Copper Targets and Copper Shaped Charges”,
Mater. Sci. Engng., in press.
11. T. Chandra (Ed.), Recrystallization ‘90, TMS, Warrendale, PA, 1990.
12. E. Nes, MetalSci., 13,211 (1979).
13. Q. Liu, M. Huang, M. Yao, and J. Yang, Acta MetaN. Mater., 40, 1753 (1992).
14. K. Tsuzaki, II. Xiaoxu, and T. Maki, Acta Mater.. 44(1 l), 4491 (1996).

You might also like