Joint Zone Evolution in Infrared Bonded Steels

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Joint Zone Evolution in Infrared Bonded Steels with

Copper Filler
J.H. LI and R.Y. LIN

Low carbon steels have been joined using an infrared processing technique with copper as the filler
material. Single lap specimens were prepared. The joining temperatures were 1100 ⬚C, 1125 ⬚C, and
1150 ⬚C with joining time ranging from 0 to 2 minutes in flowing argon. Excellent wetting between
the base materials and the filler was observed for all samples. The joint shear strength was determined
with a specially designed fixture to minimize the bending moment of specimens during testing. The
measured joint shear strength varies from 240 to 300 MPa depending on the processing conditions.
The maximum strength obtained is about 300 MPa, which can be achieved by processing at 1125 ⬚C
for 60 seconds or 1150 ⬚C for only 30 seconds. Higher processing temperatures or longer processing
time than these conditions did not improve the joint strength. Joint cross-section examinations revealed
that there are no voids in the joint. Microhardness tests performed on the cross sections of joined
samples across the joint zone indicate that the joint zone hardness is higher than that of pure copper.
Examinations with energy dispersive analysis of X-ray revealed presence of iron in the joint as well
as a small amount of copper inside the base materials.

I. INTRODUCTION reduce the inventory cost for the parts that otherwise would
have to be prepared separately and stored for assembly line
JOINING of steels for low temperature applications has needs. Other advantages include no vacuum requirement
traditionally been achieved with vacuum brazing at tempera-
even for joining superalloys or titanium alloys, controlled
tures between 1120 ⬚C and 1150 ⬚C for hours using copper
joint interfaces with limited reactions, little to no base mate-
as the filler material.[1–6] Steels joined under such processing rial property deterioration due to rapid processing, and poten-
conditions are recommended for room-temperature applica-
tials of localized heating for joining large machine parts.
tions, such as brake line hose connectors and steering wheels So far, infrared joining techniques of steels, nickel-based
on automobiles. There are millions of steel parts prepared
superalloys, titanium alloys, aluminides, silicon carbide, and
with vacuum brazing annually. Due to vacuum requirements,
titanium-matrix composites have been successfully devel-
traditionally these parts are prepared by remote vacuum
oped. In this article, infrared joining of steels with copper
brazing plants and stored in a large quantity for assembly
as the filler material was investigated.
lines. Huge inventory spaces and cost are required.
Previous studies have shown that with conventional vac-
uum brazing, due to long processing time, active penetration II. EXPERIMENTAL
of molten copper along the grain boundaries of steel in
brazing lead to the formation of cracks and failure of compo- A. Materials
nents.[1] The degree of intergranular attack of steels by mol- 1008 low carbon steel coupons of 20 ⫻ 5 ⫻ 2 mm in
ten copper increased with increasing time of processing.[2] dimension were cut and polished to 1200 mesh for joining.
Growth of columnar phase containing Fe, Cu, and C was Commercially available copper sheets (Alfa, Copper Foils,
observed in the joint. The joint strength increased with the 0.25 mm thickness, 99.9985 pct pure) were rolled to 50 ␮m
increasing amount of the columnar phase.[5,6] These studies in thickness and cut to a size of 5 ⫻ 2 mm to be used as
pointed out disadvantages of vacuum furnace brazing includ- filler coupons. Both steel and copper coupons were cleaned
ing vacuum requirements, flux requirements, long proc- with a dilute nitric acid solution, distilled water, and acetone
essing time, weakened steel parts due to the liquid copper sequentially in an ultrasonic cleaner prior to joining.
attack, and high inventory cost.
In the past ten years, a new joining technique using infra-
red was developed.[7–12] Infrared joining provides bonds that B. Infrared Joining
are superior or similar to those prepared with either the
vacuum brazing or diffusion bonding processes as reported Two steel coupons were placed in a graphite fixture
in the literature. The greatest advantages of infrared bonding machined to hold specimens in place with a lap area of about
over existing commercially available joining processes are 5 ⫻ 2 mm and a copper coupon in the lap area between the
its low cost and potential of being easily adapted to become steels. The entire specimen arrangement was held in position
a work station in the production line. This will significantly in a graphite fixture with graphite screws. A chromel-alumel
thermocouple was placed in between the graphite fixture
and the specimen immediately next to the lap area. The
specimens, with the filler metal, were then placed in an
J.H. LI, Graduate Student, and R.Y. LIN, Professor, are with the Depart-
ment of Materials Science and Engineering, University of Cincinnati, Cin- infrared furnace in a single lap joint configuration. No exter-
cinnati, OH 45221-0012. nal pressure was applied to the parts being joined. The
Manuscript submitted March 16, 2001. specimens were argon purged for approximately 30 seconds

METALLURGICAL AND MATERIALS TRANSACTIONS B VOLUME 32B, DECEMBER 2001—1177


15 seconds and a 10 g load (0.098 N) were used for the
measurement. The indent diagonals were measured with an
optical microscope at a magnification of 400 times.

III. RESULTS
A. Joint Cross Section Examinations
The cross sections for all samples processed under differ-
ent conditions were examined with an optical microscope.
Figure 2 shows the cross-section optical micrographs of
joints prepared at 1100 ⬚C, 1125 ⬚C, and 1150 ⬚C for 0, 30,
60, and 120 seconds. These micrographs show that there
were very few voids in the joint. At the opening of the
joint, copper flowed well around the steel sample, indicating
excellent wetting between the copper filler material and the
base metal at all processing temperatures and times. At a
high magnification (Figure 3), spots can be identified in the
Fig. 1—Heating cycles. A: infrared heating used in this study, B: induction
heating, and C: resistance furnace heating. joint zero. These spots have been verified with electron
microscopy as iron precipitates that had formed in the joint
upon solidification of molten copper. Due to rapid cooling,
which caused a large degree of undercooling for solidifica-
prior to the heating cycle. During the entire joining process, tion of copper and precipitation of iron, these precipitates
argon was purged through the heating chamber at 6 L/min to were small in size (less than 1 ␮m in diameter). Figure 3
prevent oxidation. Typically, the temperature of the specimen also shows that copper had protruded into the base material.
was brought to the preset joining temperature in 60 seconds Since grain boundaries are areas of high energy, it was
and, then, held at that temperature for various lengths of time anticipated that copper would have attacked the base materi-
before the power was terminated instantly. After joining, the als starting preferentially from the grain boundaries. To
sample was cooled in a flowing argon atmosphere. The determine the path of copper penetration, polished joint cross
cooling rate was rapid since the furnace wall was cold during sections were etched with a 1 pct nitric acid solution for
infrared heating, and only the sample was heated to the about 3 seconds to reveal the grain boundaries of steels. As
desired joining temperature. The samples normally cooled presented in Figure 4, penetration of copper started at points
to below 800 ⬚C in less than 10 seconds. Figure 1 illustrates of grain boundary and surface interception and extended into
the actual infrared heating cycles used in this study. For the steel body along grain boundaries. Another interesting
comparison, typical heating cycles of induction heating and observation is that the etched samples (Figure 4) show a
resistance furnace heating were also included. significantly smaller amount of particles in the joint zone
than the unetched samples (Figure 3). Since the etching
solution (1 pct nitric acid) dissolves Fe grain boundaries
C. Mechanical Testing and Joint Characterization and particle surfaces preferentially, this phenomenon also
Shear test was used to evaluate the joint strength.[9–12] indirectly indicates that the small particles in the joint zone
This method for joint shear strength determination has been are Fe precipitates.
successfully used without premature failure. The cross-head
speed used for the test was 0.254 mm/min.
B. Joint Shear Strength
Infrared joined specimens were cut through the joining
area, mounted, and polished for cross section area examina- The shear strength of the joint was determined with an
tions with an optical microscope and a scanning electron Instron testing machine (Model 4206, Instron Corp., Canton,
microscope (SEM, Hitachi, model 4000-S) powered with a MA). The specimen holder for the shear test was designed
field emission electron source. Compositions of the joining so that essentially no bending moment was induced on the
zone and the base materials near the joining zone were specimen during testing to eliminate the bending effect on
determined with an energy dispersive spectroscope (EDS, the measured joint shear strength. At least three joints were
Oxford System) equipped on the SEM. prepared for each processing condition. Figure 5 shows the
Microhardness of the joint sample cross sections was plot of measured joint shear strength as a function of the
determined by using a M-400-H1 microhardness tester from processing temperature and time. Actual strength data with
LECO. The microhardness testing procedures followed the the standard deviations calculated from all specimens were
ASTM standard test method for microhardness of materials. reported in Table I. To determine the debonding path, frac-
All microhardness tests were carried out in areas where the tured specimens after the shear test were mounted with frac-
joint zone was about 20 ␮m with the spacing between the tured surfaces facing each other, polished to reveal the joint
indents being at least 60 ␮m to eliminate possible interfer- cross sections, and examined under an optical microscope.
ences from each other. On each joint cross section, hardness It was found that all shear test fracture paths were through
measurements were done at the center of the joint, near the the joint zone, as presented in Figure 6. Thus, the measured
edge of the joint, in the base materials near the joint, and shear strength data were all for the copper filler material.
in the base materials away from the joint. An average of The joint shear strength ranges from about 240 to 300 MPa.
ten readings was taken at each location. A dwelling time of The joint strength for samples prepared at 1100 ⬚C was only

1178—VOLUME 32B, DECEMBER 2001 METALLURGICAL AND MATERIALS TRANSACTIONS B


Fig. 2—(a) through (l) Cross section optical micrographs of joints processed at 1100 ⬚C, 1125 ⬚C, and 1150 ⬚C for 0, 30, 60, and 120 s.

Fig. 3—High magnification joint zone showing spots of iron precipitates


in the joint zone. Sample prepared at 1150 ⬚C for 120 s.

up to about 260 MPa, which was achieved with 30 second


processing time. Increasing joining time to more than 30 Fig. 4—Etched joint cross section to reveal grain boundary attack of steel
by liquid Cu. Sample prepared at 1150 ⬚C for 120 s. (Etchant: 1 pct HNO3;
seconds did not increase the joint strength. Joining at 1125 and etching time: 3 s).
⬚C gave a joint strength up to 300 MPa, which was reached
with 60 seconds of joining. Further increases in the joining
time did not increase the joint strength. Joining at 1150 ⬚C
also gave a maximum joint strength of 300 MPa, which was Inspecting the joint cross section microstructure shown in
reached after 30 second joining. No further increase in the Figure 3, we can see that the joint zone contains many
joint strength was achieved when the joining time was fine precipitates, which made the joint zone a composite of
increased. For comparison, a pure copper shear specimen copper matrix reinforced with fine particles. Dissolution of
has been machined and tested. The measured shear strength Fe in copper during joining and precipitation of Fe particles
of the pure copper specimen is 174 MPa, which has the in the joint to form a composite in the joint zone are responsi-
same value as data reported in the literature.[13] It is obvious ble for the high joint strength. The higher the joining temper-
that copper in the joint zone has been strengthened. ature and the longer the joining time, the higher the Fe

METALLURGICAL AND MATERIALS TRANSACTIONS B VOLUME 32B, DECEMBER 2001—1179


Fig. 5—Joint shear strength as a function of the processing temperature
and time.

Table I. Shear Strength in (MPa) of Single Lap Specimens


from This Study
Time (s)
Temperature 0 30 60 120
1100 ⬚C 240 ⫾ 10 260 ⫾ 6 261 ⫾ 6 257 ⫾ 1
1125 ⬚C 261 ⫾ 2 271 ⫾ 2 297 ⫾ 8 296 ⫾ 6
1150 ⬚C 264 ⫾ 4 298 ⫾ 1 294 ⫾ 4 300 ⫾ 10

Fig. 7—Microhardness of joints under various processing conditions.

Fig. 6—Cross section micrograph of a shear tested joint to show the shear
fracture path. shown in the diagrams as the dotted line. This observation
agrees with results of the joint strength measurements in
that the joint zone properties are better than that of pure
concentration in molten copper until reaching saturation and, copper. The EDS analysis of the joint cross sections, reported
thus, the more precipitates exist in the joint and the stronger in Section III.D, further verified that all joint zones contain
the joint becomes. iron as a result of liquid metal attack on the base steel. The
hardness of the base steel near the joint is also higher than
that of the base steel away from the joint zone. The EDS
C. Microhardness analysis of the joint cross sections also indicates that copper
To determine the effect of joining conditions on the joint did diffuse into the base steel specimens.
and on base materials near the joint, microhardness tests
have been done on joint cross sections. Vickers diamond-
D. SEM/EDS Analysis
pyramid indenters were used in this study. Figure 7 shows
the measured microhardness of the joints under various proc- Figure 8 shows the results of SEM/EDS analysis on the
essing conditions. To minimize confusion of data presenta- joint. As mentioned previously, optical micrographs of the
tion, no error bars were included in Figure 7. Table II shows joint cross sections showed that spots existed in the joint
the average hardness values along with the standard devia- zone (Figure 3). The SEM micrograph of the polished cross
tion of the data for samples prepared at 1150 ⬚C. The standard section of the joint (Figure 8(a)) was not able to reveal those
deviations for hardness data for 1100 ⬚C and 1125 ⬚C were spots in the joint. However, EDS elemental mapping of iron
similar to those for 1150 ⬚C samples. For all joints, the joint on the same area indeed identified iron-rich spots in the
zone hardness is higher than that of pure copper, which is joint zone, as shown by circles in Figure 8(b). The size of

1180—VOLUME 32B, DECEMBER 2001 METALLURGICAL AND MATERIALS TRANSACTIONS B


Table II. Microhardness at the Joint Cross Sections (at 1150 ⴗC)

Processing Distance from Joint Center (␮m)


Time (s) 0 10 20 30 40 50
0 72.2 ⫾ 0.5 105.6 ⫾ 4.3 116.8 ⫾ 4.8 109.6 ⫾ 5.2 102.6 ⫾ 1.5 104.4 ⫾ 3.4
30 76.6 ⫾ 0.7 109.5 ⫾ 5.9 117.4 ⫾ 8.2 108.0 ⫾ 3.2 105.1 ⫾ 4.8 102.0 ⫾ 1.6
60 74.0 ⫾ 1.8 103.4 ⫾ 2.1 119.3 ⫾ 1.4 110.0 ⫾ 1.8 101.8 ⫾ 1.5 104.6 ⫾ 6.5
90 76.1 ⫾ 1.8 110.6 ⫾ 2.6 119.2 ⫾ 3.3 113.3 ⫾ 4.4 106.8 ⫾ 2.7 104.1 ⫾ 4.8

Table III. EDS Results at Various Locations in the Joint


Shown in Figure 8(a)
Atomic Percent
Position Fe (K␣) Cu (K␣)
1 98 2
2 99 1
3 99 1
4 43 57
5 6 94
6 5 95
7 6 94

these spots is less than 1 ␮m in diameter. Upon rapid cooling


at the end of infrared joining, a large degree of undercooling
had occurred for solidification of copper and precipitation
(a) SEM photo of iron. As a result, the iron precipitates were small in size
(less than 1 ␮m in diameter). Table III lists EDS analysis
data at several locations across the joint. The overall amount
of Fe in the joint agrees with that predicted by the phase
diagram of the Cu-Fe binary system.[14] Point 4 on the SEM
micrograph in Figure 8(a) was located at the interface
between the base material and the joint. The electron spot
size for EDS analysis was typically between 3 and 5 ␮m.
Therefore, the EDS results at this point reflect influences
from both the base and the joint materials. Cu mapping with
EDS on the same area is shown in Figure 8(c). Four features
stand out from copper mapping. First, the joint zone is
primarily copper, and some copper was found in the base
material. Second, the intensity of Cu at the spots of iron in
the joint zone is lower than that around the spots verifying
that these spots indeed are iron precipitates. Third, at the
(b) Fe EDS mapping joint-base material interfaces, copper protruded into the base
material. This is in agreement with that observed in Figures
3 and 4 in that copper had dissolved iron preferentially at
the grain boundaries. Finally, the Cu concentration along
the grain boundaries in the base material is greater than that
in the grains of the base material. Slightly higher Cu intensity
along the grain boundaries than that in the grains can be
identified in the copper mapping diagram, as illustrated by
the arrows in Figure 8(c). This observation suggests that
there was preferential diffusion of copper in the grain bound-
aries of steel used in this study.

IV. DISCUSSION
A. Joint Evolution
Evidences from joint microstructures, the joint strength,
(c) Cu EDS mapping and hardness measurements, as well as SEM/EDS analysis,
Fig. 8—(a) through (c) SEM micrograph and EDS mapping of a joint suggest that the following processes had occurred during
cross section. joining. Due to rapid infrared heating, copper in the lap area

METALLURGICAL AND MATERIALS TRANSACTIONS B VOLUME 32B, DECEMBER 2001—1181


steels was through both the grain boundaries and the steel
grains. The grain boundary diffusion proceeds much faster
than that through the grains. Since there is no intermetallic
compound in the Cu-Fe binary system, no reaction product
will be presented in the joint or in the base materials where
copper had diffused to.
At the end of joining, as soon as the infrared power was
shut off, the specimen temperature would immediately drop
to below the copper melting point, 1083 ⬚C, in less than 10
seconds. Liquid copper with Fe in the solution would solid-
ify. Since the solubility of Fe in solid copper is lower than
that in liquid copper, Figure 9, Fe particles would precipitate
out in the joint. The final joint area will be copper with fine
Fe precipitates distributed throughout the joint, as presented
in Figure 3. Figure 9 also shows that the amount of Cu
that can be dissolved in the base material at the joining
Fig. 9—Cu-Fe phase diagram with regions of experimental conditions in temperature is higher than the solubility limit of Cu in Fe
this study. at room temperature. However, due to rapid cooling, no Cu
will nucleate out in the base materials when the sample was
cooled to room temperature. Thus, there will be supersatu-
rated Cu in the base materials near the joint. The higher
hardness values in the base material next to the joint than
that far away from the joint confirm such supersaturation
arguments. Penetration of Cu into steel is much deeper than
that predicted by the lattice diffusion of Cu in Fe since grain
boundaries of steels provide an easy path for Cu penetration.
The joining mechanism presented previously also agreed
with observations obtained previously by other researchers[1–6]
in terms of dissolution of Fe in molten copper and inter-
granular attack of Fe by molten copper. From this study, we
further observed that Cu diffuse into Fe preferentially
through grain boundaries resulting in deep Cu penetration
in steels. Formation of copper matrix composites reinforced
with fine Fe particles was also observed in this study. The
high joint strength obtained in this study is due to the com-
posite formation.

Fig. 10—The Cu rich end of the Cu-Fe phase diagram. B. Joint Evolutions vs Joint Strength, Hardness Data,
and EDS Analysis Results
The hardness data presented in Figure 7 show that the
quickly melted shortly after turning the power on. In contact joint center hardness is higher than that of pure copper. The
with liquid copper, the steel surfaces in the lap area would higher the joining temperature, the higher the joint center
dissolve. The longer the joining time, the more dissolution hardness. Since the solubility of Fe in Cu increases with
of steel in liquid copper would be until reaching saturation. increasing temperature and time, the more Fe dissolution in
Also, the higher the joining temperature was the higher the liquid copper during joining, the more precipitates in the
solubility of Fe in copper would be as indicated in the Cu- Cu joint presented and, thus, the higher the expected hard-
Fe phase diagram, Figure 9, and the enlarged Cu rich end ness. The hardness measurements support the joint evolution
of the phase diagram in Figure 10. The attack of liquid copper process described previously.
on steel occurred preferentially on the grain boundaries, as In the base material, the hardness measurements show
illustrated in the joint microstructures (Figures 3 and 4). that near the joint the microhardness has the highest value
The Cu-Fe phase diagram also indicates that at the joining and levels off to the hardness of typical 1008 steel away
temperatures, the stable iron phase is ␥ -Fe with a fcc struc- from the joint. Although the precision of the microhardness
ture and, at room temperature, Fe has a bcc structure. How- measurement may not be sufficient to distinguish the differ-
ever, it was assumed that the steel will maintain its bcc ence in hardness data as affected by limited differences in
structure at the joining temperatures because the joining time the amounts of copper supersaturation in steels, the general
was less than 120 seconds. trend of the hardness data variations agree with the joint
In the meantime, copper would have also diffused into evolution process presented previously. Examinations of the
steel following a solid state diffusion process. The concentra- hardness data near the joint show that the base material
tion of Cu in Fe immediately next to the copper zone could hardness reaches the highest value of around 119 HV when
be very close to the solubility limit of Cu in metastable ␣Fe joined at 1150 ⬚C and has the lowest hardness at around
extrapolated from the low temperature region to the joining 109 HV when joined at 1100 ⬚C.
temperatures, as shown in Figure 9. Diffusion of copper in Elemental mapping with EDS on the cross sections of

1182—VOLUME 32B, DECEMBER 2001 METALLURGICAL AND MATERIALS TRANSACTIONS B


joined samples revealed iron rich spots in the joint and X-ray revealed presence of iron in the joint and diffusion
verified the iron precipitation process during joining sug- of copper into the base materials.
gested previously. The higher Cu concentration along the 5. Cu diffuses into Fe preferentially through grain bound-
grain boundaries of steels than that in the grains after joining aries resulting in Cu penetration in steels.
provided evidences of preferred grain boundary diffusion. 6. Formation of copper matrix composites reinforced with
Such preferred diffusion paths along grain boundaries allow fine Fe particles was observed in the joint zone. The
deep penetration of Cu into steel. Microhardness measure- high joint strength obtained in this study is due to the
ments of the joint cross sections showed that, in the steel, composite formation.
regions with higher hardness than that of the original steel
due to alloying by copper atoms are deeper than the distance
of copper migration calculated from simple lattice diffusion ACKNOWLEDGMENT
of Cu in steels. Such a large diffusion distance is attributed to The authors are thankful for the support of the WSU-
the fast diffusion path of Cu through steel grain boundaries. PCC program. Valuable discussion and input from graduate
students in the same group, J.W. Seok, K. Chokalingam,
P. Deshpande, Y. Wang, and H. Mu are also very much
V. CONCLUSIONS appreciated.
The following conclusions have been obtained from
this study. REFERENCES
1. V.N. Radzievskii, Y.F. Gartsunov, L.V. Barnov, and V.M. Rab: Welding
1. Low carbon steels have been successfully joined using Int., 1991, vol. 5 (2), pp. 148-49.
an infrared processing technique with copper as the filler 2. W.F. Savage, E.F., Nippes, and R.P. Station: Welding J., 1978, vol.
material. The joining temperatures were 1100 ⬚C, 1125 57 (1), pp. 9s-16s.
⬚C, and 1150 ⬚C with joining time ranging from 0 to 2 3. W.D. Kehr: Met. Progr., 1966, vol. 89 (1), pp. 90-93.
4. I. Kawakatsu and T. Osawa: Welding J., 1977, vol. 56 (2), pp. 56S-60S.
minutes in flowing argon. 5. T. Yoshida and H. Ohmura: Welding J., 1980, vol. 59 (10), pp.
2. Excellent wetting between the base materials and the 278s-282s.
filler was observed for all samples. There were no voids 6. T. Yoshida and H. Ohmura: Welding J., 1985, vol. 64 (1), pp. 1s-12s.
in the joint when the strength reached its maximum value. 7. S.J. Lee, S.K. Wu, and R.Y. Lin: Acta Mater., 1998, vol. 46 (4),
pp. 1283-96.
3. The measured compressive shear strength varies between 8. C.A. Blue, V.K. Sikka, R.A. Blue, and R.Y. Lin: Metall. Mater. Trans.
240 and 300 MPa depending on the processing condi- A, 1996, vol. 27A, pp. 4011-18.
tions. The maximum strength obtained from infrared 9. C.A. Blue, R.A. Blue, and R.Y. Lin: Scripta Metall. Mater., 1995, vol.
processing is around 300 MPa, which can be achieved 32 (1), pp. 127-32.
by processing the specimens at 1125 ⬚C for 60 seconds 10. R.Y. Lin, S.G. Warrier, C.A. Blue, C.C. Chen, C.A. Eppich, and
R.A. Blue: JOM, 1994, vol. 46 (3), pp. 26-30.
or 1150 ⬚C for only 30 seconds. Higher processing tem- 11. C.A. Blue and R.Y. Lin: MRS Symp. Proc., 1993, vol. 314, pp. 143-48.
peratures or longer processing time than these conditions 12. C.A. Blue, S.G. Warrier, M.T. Robson, and R.Y. Lin: Welding J., 1993,
did not improve the joint strength. vol. 72 (6), pp. 51-54.
4. Microhardness tests across the joint zone on polished 13. C.J. Smithells: Metals Reference Book, ASM INTERNATIONAL,
Materials Park, OH, 1992, pp. 13.9-13.79.
cross sections indicate that the microhardness in the joint 14. M. Venkatraman and J.P. Newmann: Binary Alloy Phase Diagrams,
is greater than that of the original copper filler material. 2nd ed., ASM INTERNATIONAL, Materials Park, OH, 1990, pp.
Elemental analysis by energy dispersive analysis with 1408-10.

METALLURGICAL AND MATERIALS TRANSACTIONS B VOLUME 32B, DECEMBER 2001—1183

You might also like