You are on page 1of 8

j m a t e r r e s t e c h n o l .

2 0 1 4;3(2):114121

Available online at www.sciencedirect.com

www.jmrt.com.br

Original Article

A metalmetal bonding process using metallic copper


nanoparticles produced by reduction of copper oxide
nanoparticles
Yoshio Kobayashi a,, Yuki Abe a, Takafumi Maeda a, Yusuke Yasuda b , Toshiaki Morita b
a
b

Department of Biomolecular Functional Engineering, College of Engineering, Ibaraki University, Hitachi, Ibaraki, Japan
Hitachi Research Laboratory, Hitachi Ltd., Hitachi, Ibaraki, Japan

a r t i c l e

i n f o

a b s t r a c t

Article history:

Metalmetal bonding was performed using metallic Cu nanoparticles fabricated from CuO

Received 20 September 2013

nanoparticles. Colloid solutions of CuO nanoparticles were prepared by the reaction of

Accepted 24 December 2013

Cu(NO3 )2 and NaOH in aqueous solution at reaction temperatures (TCuO s) of 20, 40, 60 and

Available online 6 February 2014

80 C. CuO single crystallites with a size of ca. 10 nm were produced, and they formed leaflike aggregates. The longitudinal and lateral sizes of the aggregates decreased from 522 to

Keywords:

84 nm and from 406 to 23 nm, respectively, with an increase in TCuO . Colloid solutions of

Metallic Cu

metallic Cu nanoparticles were prepared by reducing the CuO nanoparticles with hydrazine

Copper oxide

in the presence of cetyltrimethylammonium bromide. The size of the metallic Cu nanopar-

Nanoparticle

ticles decreased from 92 to 73 nm with increasing TCuO . Metallic copper discs were bonded

Metalmetal bonding

with the metallic Cu nanoparticles by annealing at 400 C and 1.2 MPa for 5 min in H2 gas.
The shear strength required to separate the bonded discs increased with increasing TCuO . A
maximum shear strength of 39.2 MPa was recorded for a TCuO of 80 C.
2014 Brazilian Metallurgical, Materials and Mining Association. Published by Elsevier
Editora Ltda. All rights reserved.

1.

Introduction

In metalmetal bonding processes, which are important


in many elds, solders or llers have conventionally been
used for efcient bonding [15]. These solders are melted
at high temperatures and spread between metallic surfaces,
thus bonding the surfaces together. A decrease in temperature solidies the metallic materials and completes the
metalmetal bonding. Metallic alloys composed mainly of lead
and tin have been used as solders [14]. These metallic alloys

melt at temperatures as low as 184 C, lower than the melting


points of many other metallic alloys. The lead- and tin-based
alloys diffuse into the materials to be bonded, and then, they
can be bonded at low temperatures. It is well known that
lead is harmful to living bodies, which limits its use. Various
lead-free alloys have been developed as new solders [611].
Although low-temperature metalmetal bonding can be performed using the lead-free solders, there is a serious problem:
the bonded materials may break apart when exposed to temperatures higher than their melting points due to remelting of
the solders.

Corresponding author.
E-mail: ykoba@mx.ibaraki.ac.jp (Y. Kobayashi).

2238-7854/$ see front matter 2014 Brazilian Metallurgical, Materials and Mining Association. Published by Elsevier Editora Ltda. All rights reserved.

http://dx.doi.org/10.1016/j.jmrt.2013.12.003

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

Metallic materials such as Au, Ag and Cu can be used as


the ller because they have excellent electrical and thermal
conductivities. However, their melting points are ca. 1000 C,
higher than those of the conventional lead- and tin-based
solders. High-temperature annealing is required during the
bonding process to successfully bond metallic materials, and
these high temperatures thermally damage the material near
the bonding site.
The melting points of metallic materials such as Au, Ag and
Cu are ca. 1000 C in the bulk state but decrease as the material
size is decreased to several nanometers [1216]. This decrease
in the melting point decreases the temperature needed for
the metalmetal bonding process. Once the metallic materials are bonded with the metallic nanoparticles, they remain
bonded even at temperatures higher than the melting points
of the metallic nanoparticles because the nanoparticles convert to a bulk state during bonding. Various researchers have
studied metalmetal bonding using metallic Ag nanoparticles as the ller [15,1723] because metallic Ag is chemically
stable. Although Ag nanoparticles work well as a ller for
metalmetal bonding, they have some disadvantages: metallic Ag is relatively expensive and prone to migration under an
applied voltage, which may damage an electric circuit.
Metallic Cu is promising as a ller for bonding because it
is inexpensive and because electric migration does not take
place as often as it does with metallic Ag. Several researchers
have studied metalmetal bonding using metallic Cu nanoparticles [20,24,25]. Our research group has developed a method
for producing metallic Cu nanoparticles for use as the ller
in metalmetal bonding [2630]. Metallic Cu nanoparticles are
usually prepared using methods based on the direct reduction of copper ions with reducing reagents in the liquid phase.
Copper oxide can also be used as a precursor. Previously, our
research group proposed a method for fabricating metallic Cu
nanoparticles using CuO nanoparticles as the precursor and
examined the metallic Cu nanoparticles as a ller for bonding
[31]. We also varied the morphologies of CuO nanoparticles
and CuO-nanoparticle aggregates with several parameters
such as the reaction temperature [32]. The morphology of
metallic Cu particles depended on the morphologies of the
CuO nanoparticles and CuO-nanoparticle aggregates as well
as on the concentrations of chemicals. The metalmetal bonding properties are related to the morphology of the metallic
Cu nanoparticles. In the present work, metallic Cu nanoparticles are fabricated from CuO nanoparticles with various
preparation parameters and are examined with respect to
metalmetal bonding.

2.

Experimental work

2.1.

Chemicals

The reactants for the production of copper oxide were copper (II) nitrate trihydrate (Cu(NO3 )2 3H2 O) (77.080.0% (as
Cu(NO3 )2 )) and sodium hydroxide solution (NaOH) (5 M).
Hydrazine monohydrate (>98.0%) and cetyltrimethylammonium bromide (CTAB) (99%) were used as a reducing reagent
for copper oxide and as a stabilizer for the resulting metallic
Cu nanoparticles, respectively. All chemicals were purchased

115

from Kanto Chemical Co., Inc. and were used as received.


Water that was ion-exchanged and distilled with a Yamato
WG-250 system was used in all preparations, and it was deaerated by bubbling with N2 gas for 30 min prior to preparation
of the aqueous solutions of Cu(NO3 )2 .

2.2.

Preparation

Colloid solutions of copper oxide nanoparticles were synthesized using a reaction between copper ions and a base, as
described in our previous work [31]. An aqueous solution of
NaOH was added to an aqueous solution of Cu(NO3 )2 with
vigorous stirring at reaction temperatures (TCuO s) of 20, 40, 60
and 80 C. The reaction times were 3 h at TCuO s of 60 and 80 C
and 24 h at TCuO s of 20 and 40 C. The initial concentrations
of Cu(NO3 )2 and NaOH were 1.25 102 and 2.375 102 M,
respectively, which resulted in an Na/Cu molar ratio of 1.9 in
the nal solution.
Colloid solutions of metallic Cu nanoparticles were synthesized by reducing the copper oxide nanoparticles. An aqueous
solution of CTAB was added to the colloid solution of copper
oxide nanoparticles. After 30 min, hydrazine was added. Both
the additions were performed in air with vigorous stirring at a
constant solution temperature of 30 C, and the reaction time
was 3 h. Initial concentrations of Cu, hydrazine and CTAB in
the nal colloid solution were 1.0 102 , 0.8 and 5.0 103 M,
respectively. The resulting nanoparticles were washed rst
with centrifugation at 10,000 rpm for 30 min, followed by the
removal of the supernatant, the addition of water, and shaking
of the mixture with a vortex mixer to disperse the nanoparticles. This process was performed three times. The powdered
nanoparticles were obtained by drying the nanoparticles at
room temperature in a vacuum after the nal removal of the
supernatant.

2.3.

Characterization

The particles were characterized by transmission electron


microscopy (TEM) and X-ray diffractometry (XRD). TEM observation was performed with a JEOL JEM-2100 microscope
operating at 200 kV. Samples for TEM were prepared by dropping and evaporating the nanoparticle colloid solution on a
collodion-coated copper grid. Dozens of particle diameters
were measured in TEM images to determine an average particle size and a standard deviation. XRD patterns were obtained
with a Rigaku Ultima IV X-ray diffractometer at 40 kV and
30 mA with Cu K1 radiation. Samples for XRD were prepared by removing the supernatant of the nanoparticle colloid
solution with decantation and drying the residue at room temperature for 24 h in a vacuum.
The metalmetal bonding properties of metallic Cu
nanoparticles were investigated by a method described in our
previous work [21,3335]. The powdered particles, which were
prepared using the same process as that used for the XRD samples, was spread on a copper stage with a diameter of 10 mm
and a thickness of 5 mm. A copper plate with a diameter of
5 mm and a thickness of 2.5 mm was placed on top of the
powder sample. The copper stage and plate were pressed at
1.2 MPa while annealing in H2 at 400 C for 5 min with a Shinko
Seiki vacuum reow system. After bonding, the copper stage

116

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

and plate were separated by applying a shear strength, which


was measured with a Seishin SS-100KP bond tester. The bonding and the measurement of shear strength were performed
two or three times for each sample and averaged to determine
the bonding strength of the nanoparticles. The surface of each
copper plate was observed by scanning electron microscopy
(SEM) with a JEOL JSM-5600LV microscope after the measurement of shear strength.

3.

Results and discussion

3.1.

Morphology of copper oxide nanoparticles

The blue aqueous solution of Cu(NO3 )2 became a black colloid solution after the addition of the NaOH solution. Because
CuO is black, this observation implied the production of CuO.
Fig. 1 shows the TEM images of the prepared nanoparticles, of
which colloid solutions were prepared at various TCuO s. At a
TCuO of 20 C, leaf-like aggregates with an average longitudinal
size of 522 nm and an average lateral size of 406 nm were produced. A high magnication image (inset of Fig. 1(a)) revealed
that the aggregates appeared to be composed of nanoparticles with a size of ca. 10 nm, which is only an estimate, as
the outlines of the nanoparticles were not well dened in the
TEM images. The addition of NaOH made the pH of the solution approach its isoelectric point (IEP) without exceeding it
because the Na/Cu ratio of 1.9 was lower than the stoichiometric ratio of 2 required for complete reaction of NaOH and

Cu2+ . This increase in pH decreased the surface charge of


the nanoparticles, which resulted in their aggregation. The
aggregate size decreased with increasing TCuO ; the longitudinal/lateral aggregate sizes (in nm) were 522 43/406 39
at 20 C, 208 22/151 9 at 40 C, 75 10/48 9 at 60 C and
84 15/23 5 at 80 C. This trend corresponded with our previous work [32], which examined the fabrication of CuO
particles at initial concentrations of 1.0 102 M Cu(NO3 )2
and 1.9 102 M NaOH. In our previous work [31], the IEP
of CuO particles prepared at 80 C was higher than that of
CuO nanoparticles prepared at 20 C. Therefore, at a TCuO
of 80 C, the pH did not closely approach the IEP. Consequently, electrostatic repulsions between the nanoparticles
limited aggregation, and small aggregates were produced at
high TCuO s.
Fig. 2 shows the XRD patterns of the CuO nanoparticles.
For the particles prepared at 20 C, the dominant peaks were
at 12.8 and 25.8 . These peaks were assigned to Cu2 (OH)3 NO3
(JCPDS card No. 15-0014). In addition, several peaks attributed
to monoclinic CuO (JCPDS card No. 5-0661) were recorded
at 35.6 , 38.8 , and 48.9 . The appearance of Cu2 (OH)3 NO3
peaks indicated that the TCuO of 20 C was too low for complete reaction of Cu(NO3 )2 to CuO. At TCuO s over 20 C, the
Cu2 (OH)3 NO3 peaks disappeared, and only the peaks due to
CuO were detected. Accordingly, it was found that the purity
of CuO nanoparticles increased with increasing TCuO . Application of the Scherrer equation to the XRD line broadening of
the 35.6 peak provided average CuO crystal sizes of 12.6, 14.3,
13.4 and 10.7 nm for 20, 40, 60 and 80 C, respectively. There

Fig. 1 TEM images of various nanoparticles. Colloid solutions of samples (a)(d) were prepared with the saltbase reaction
at 20, 40, 60, and 80 C.

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

was no large difference in crystal size at the TCuO s examined.


These crystal sizes of ca. 10 nm were similar to the sizes of
the nanoparticles composing the aggregates. This similarity
indicated that single crystals of CuO formed the aggregates.

Intensity (arb. units)

3.2.

a
10

20

117

30

40

50

60

70

80

2 (degree), CuK
Fig. 2 XRD patterns of various nanoparticles. Samples
(a)(d) were the same as in Fig. 1. Symbols () and () stand
for CuO and Cu2 (OH)3 NO3 , respectively.

Morphology of metallic copper nanoparticles

The addition of hydrazine reddened the colloid solution of CuO


nanoparticles. Since surface plasmon resonance absorption
of metallic Cu nanoparticles results in a red color, the reddening implied the production of metallic Cu nanoparticles.
Fig. 3 shows TEM images of the metallic Cu particles. The particles had rough surfaces. Although some particles larger than
1 m were observed, almost all of the nanoparticles examined
had sizes in the range of 50150 nm. The average particle sizes
were 92 33, 93 33, 101 37 and 73 23 nm for TCuO s of 20,
40, 60 and 80 C, respectively; the particle size was approximately constant in the TCuO range of 2060 C but decreased
when the TCuO was increased to 80 C. As shown in Fig. 1, the
aggregate size decreased with increasing TCuO . Accordingly,
aggregate size was considered a reection of particle size.
Fig. 4 shows the XRD patterns of the metallic Cu nanoparticles. The patterns showed peaks at 43.3 , 50.3 and 74.2 . These
peaks were attributed to cubic metallic Cu (JCPDS card No.
4-0836). In addition to the metallic Cu peaks, a faint peak was
detected at 36.6 in each pattern. This peak was assigned to
cubic metallic Cu2 O (JCPDS card No. 5-0667). The detection of
Cu2 O indicated that the nanoparticles were partially oxidized.
The average crystal sizes of the metallic Cu nanoparticles,

Fig. 3 TEM images of various nanoparticles. Colloid solutions of samples (a)(d) were prepared by reducing the samples
(a)(d) in Fig. 1, respectively.

118

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

size, which indicated that the nanoparticles were polycrystalline.

3.3.
Bonding properties of metallic copper
nanoparticles
Intensity (arb. units)

a
10

20

30

40

50

60

70

80

2 (degree), CuK
Fig. 4 XRD patterns of various nanoparticles. Samples
(a)(d) were the same as in Fig. 3. Symbols () and () stand
for metallic Cu and Cu2 O, respectively.
which were estimated from the XRD line broadening of the
43.3 peak using the Scherrer equation, were 33.1, 31.1, 28.2
and 32.6 nm at TCuO s of 20, 40, 60 and 80 C, respectively. The
particle size observed with TEM was larger than the crystal

Fig. 5 shows the photographs of the copper plates after the


measurement of shear strength. For all the particles examined, the pressed powders on the plates appeared to be faintly
reddish due to the presence of metallic Cu: the powders
were still metallic after the bonding process. This observation implies that the metallic Cu powders strongly bonded the
copper discs together.
Fig. 6 shows shear strength as a function of TCuO . All of the
shear strengths measured were over 20 MPa, which indicated
that the copper discs were strongly bonded for all the particles examined and also corresponded to the strong bonding
implied in Fig. 5. The shear strength increased with increasing TCuO and reached 39.2 MPa at a TCuO of 80 C, which was the
largest in the present work. As shown in Fig. 3, the particle size
decreased with increasing TCuO . Small particles have a larger
surface area than large particles if the amount of material
is the same. Therefore, the contact area between the particles and the Cu discs was larger for the small particles, as
a result of the larger surface area, than for the large particles. This large contact area for the small particles probably
caused the strong metalmetal bonding. There are other possible causes as well. One possible cause involves impurities
in the metallic Cu nanoparticles. Our previous work showed

Fig. 5 Photographs of the copper plates after the measurement of shear strength. The powders used were the metallic Cu
particles (a)(d) shown in Fig. 3.

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

50

Shear strength (MPa)

40

30

20

10

0
20

40

60

80

TCuO (C)
Fig. 6 Shear strength vs. TCuO . The powders used were the
metallic Cu particles (a)(d) shown in Fig. 3.

that the presence of Cu2 (OH)3 NO3 decreased the bonding ability [31]: during bonding, the elimination of NO3 and H2 O
from Cu2 (OH)3 NO3 produced pores that prevented the particles from sintering, thus decreasing the shear strength. In
the present work, the CuO nanoparticles fabricated at the

119

low TCuO contained Cu2 (OH)3 NO3 , while the purity of CuO
nanoparticles increased with increasing TCuO , as indicated by
the XRD measurements (Fig. 2). Because the CuO nanoparticles were the precursor of the metallic Cu particles, the purity
of the metallic Cu nanoparticles should depend on that of the
CuO nanoparticles. The purity of the metallic Cu nanoparticles might increase with increasing TCuO , while Cu2 (OH)3 NO3
might remain in the metallic Cu nanoparticles derived from
the CuO nanoparticles produced at the low TCuO . Therefore,
the metallic Cu nanoparticles derived from the CuO nanoparticles produced at the high TCuO probably did not contain much
Cu2 (OH)3 NO3 , and elimination of NO3 and H2 O did not occur
during bonding in the metallic Cu nanoparticles at high TCuO .
Thus, no decrease in bonding ability as a result of impurities
was expected for the high TCuO sample. The other possible
cause of strong metalmetal bonding in the high TCuO sample
involves a size effect. Sintering of particles takes place intensively for small particles because of the size effect related to
the depression of the melting point. This intensive sintering
might have contributed to the strong metalmetal bonding
with high TCuO , at which the smaller nanoparticles were produced. These three factors are the possible mechanisms for
the strong bonding achieved by the nanoparticles fabricated
at the TCuO of 80 C.
Fig. 7 shows the SEM images of the surfaces of the copper
discs separated with shear stress. Some dimples accompanied
by sharp tips were observed on the surfaces for the samples
at TCuO s of 20 and 80 C. When the metallic materials are separated with shear stress, dimples often form in the bonded

Fig. 7 SEM images of the copper plates after the measurement of shear strength. The powders used were the metallic Cu
particles (a)(d) shown in Fig. 3.

120

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

region. Such dimples were produced in materials that underwent a strong bonding process using Ag nanoparticles, as
reported by Morisada et al. [20] Therefore, the observation of
dimples on the copper disks suggests that they were strongly
bonded using the metallic Cu nanoparticles. For the TCuO s of
40 and 60 C, clear dimples were not observed, but the bonded
materials appeared to be sharply cut off. This sharp cut-off
also indicated strong bonding.

4.

Conclusions

Colloid solutions of leaf-like aggregates composed of CuO single crystallites with a size of ca. 10 nm were prepared by
reacting 1.25 102 M Cu(NO3 )2 with NaOH in aqueous solution at an Na/Cu ratio of 1.9 and at TCuO s of 20, 40, 60 and 80 C.
The aggregate size decreased from 522 to 84 nm for longitudinal size and from 406 to 23 nm for lateral size with an increase
in TCuO from 20 to 80 C. The CuO nanoparticles, with a Cu
concentration of 1.25 102 M in the colloid solution, were
reduced with 0.8 M hydrazine in the presence of 5.0 103 M
CTAB in water at 30 C to produce the metallic Cu nanoparticles. The size of the metallic Cu nanoparticles decreased from
92 to 73 nm with an increase in TCuO from 20 to 80 C. The
bonding strength of the particles was examined by putting the
metallic Cu nanoparticles between two metallic copper discs
and then annealing the discs at 400 C at a pressure of 1.2 MPa
for 5 min in H2 gas. The shear strength required to separate
the bonded discs increased with increasing TCuO , reaching a
maximum of 39.2 MPa at a TCuO of 80 C. Further studies on
the practical use of metallic Cu nanoparticles and the precise
bonding mechanism are in progress.

[6]

[7]

[8]

[9]

[10]

[11]

[12]

[13]

[14]

[15]

[16]

[17]

Conicts of interest
The authors declare no conicts of interest.

[18]

Acknowledgements

[19]

This work was partially supported by Hitachi, Ltd. We express


our thanks to Prof. T. Noguchi at the College of Science of
Ibaraki University, Japan for assistance with the TEM observations.

[20]

references

[21]

[22]
[1] Tu KN, Zeng K. Tinlead (SnPb) solder reaction in ip chip
technology. Mater Sci Technol 2001;34:158.
[2] Nayeb-Hashemi H, Yang P. Mixed mode I/II fracture and
fatigue crack growth along 63Sn37Pb solder/brass interface.
Int J Fatigue 2001;23:32535.
[3] Sharif A, Chan YC, Islam RA. Effect of volume in interfacial
reaction between eutectic SnPb solder and Cu metallization
in microelectronic packaging. Mater Sci Eng B 2004;106:1205.
[4] He M, Chen Z, Qi G. Solid state interfacial reaction of
Sn37Pb and Sn3.5Ag solders with NiP under bump
metallization. Acta Mater 2004;52:204756.
[5] Lugscheider E, Bobzin K, Maes M, Ferrara S, Erdle A.
Advancement in low melting solder deposition by pulsed

[23]

[24]

[25]

Magnetron Sputter-PVD process for microsystem technology.


Surf Coat Technol 2005;200:4447.
Hindler M, Guo Z, Mikula A. Lead-free solder alloys:
thermodynamic properties of the (Au + Sb + Sn) and the
(Au + Sb) system. J Chem Thermodyn 2012;55:1029.
Fallahi H, Nurulakmal MS, Arezodar AF, Abdullah J. Effect of
iron and indium on IMC formation and mechanical
properties of lead-free solder. Mater Sci Eng A 2012;553:2231.
El-Daly AA, El-Taher AM. Evolution of thermal property and
creep resistance of Ni and Zn-doped Sn2.0Ag0.5Cu
lead-free solders. Mater Design 2013;51:78996.
Hu X, Li K, Min Z. Microstructure evolution and mechanical
properties of Sn0.7Cu0.7Bi lead-free solders produced by
directional solidication. J Alloys Compd 2013;566:23945.
Maleki M, Cugnoni J, Botsis J. Microstructure-based modeling
of the ageing effect on the deformation behavior of the
eutectic micro-constituent in SnAgCu lead-free solder. Acta
Mater 2013;61:10314.
Osrio WR, Peixoto LC, Garcia LR, Mangelinck-Nol N, Garcia
A. Microstructure and mechanical properties of SnBi, SnAg
and SnZn lead-free solder alloys. J Alloys Compd
2013;572:97106.
Shibuta Y, Suzuki T. Melting and solidication point of
fcc-metal nanoparticles with respect to particle size: a
molecular dynamics study. Chem Phys Lett 2010;498:3237.
Li H, Han PD, Zhang XB, Li M. Size-dependent melting point
of nanoparticles based on bond number calculation. Mater
Chem Phys 2013;137:100711.
Xu SH, Fei GT, Zhang Y, Li XF, Jin Z, Zhang LD.
Size-dependent melting behavior of indium nanowires. Phys
Lett A 2011;375:174650.
Eluri R, Paul BK. Silver nanoparticle-assisted diffusion
brazing of 3003 Al alloy for microchannel applications. Mater
Design 2012;36:1323.
Omar MS. Models for mean bonding length, melting point
and lattice thermal expansion of nanoparticle materials.
Mater Res Bull 2012;47:351822.
Ide E, Angata S, Hirose A, Kobayashi KF. Metalmetal bonding
process using Ag metallo-organic nanoparticles. Acta Mater
2005;53:238593.
Bai JG, Zhang ZZ, Calata JN, Lu GQ. Low-temperature
sintered nanoscale silver as a novel semiconductor
device-metallized substrate interconnect material. IEEE
Trans Compon Pack Technol 2006;29:58993.
Gain AK, Chan YC, Sharif A, Wong NB, Yung WKC. Interfacial
microstructure and shear strength of Ag nano particle doped
Sn9Zn solder in ball grid array packages. Microelectron
Reliab 2009;49:74653.
Morisada Y, Nagaoka T, Fukusumi M, Kashiwagi Y,
Yamamoto M, Nakamoto M. A low-temperature bonding
process using mixed CuAg nanoparticles. J Electron Mater
2010;39:12838.
Yasuda Y, Ide E, Morita T. Low-temperature bonding using
silver nanoparticles stabilized by short-chain alkylamines.
Jpn J Appl Phys 2009;48:125004.
Suganuma K, Sakamoto S, Kagami N, Wakuda D, Kim KS,
Nogi M. Low-temperature low-pressure die attach with
hybrid silver particle paste. Microelectron Reliab
2012;52:37580.
Oestreicher A, Rhrich T, Wilden J, Lerch M, Jakob A, Lang H.
An innovative method for joining materials at low
temperature using silver (nano) particles derived from
[AgO2 C(CH2 OCH2 )3 H]. Appl Surf Sci 2013;265:23944.
Nishikawa H, Hirano T, Takemoto T, Terada N. Effects of
joining conditions on joint strength of Cu/Cu joint using Cu
nanoparticle paste. Open Surf Sci J 2011;3:604.
Yan J, Zou G, Wu A, Ren J, Hu A, Zhou YN. Polymer-protected
CuAg mixed NPs for low-temperature bonding application. J
Electron Mater 2012;41:188692.

j m a t e r r e s t e c h n o l . 2 0 1 4;3(2):114121

[26] Kobayashi Y, Ishida S, Ihara K, Yasuda Y, Morita T.


Metalmetal bonding process using copper nanoparticles.
World J Eng 2010;7:5834.
[27] Kobayashi Y, Shirochi T, Yasuda Y, Morita T. Preparation of
metallic copper nanoparticles in aqueous solution and their
bonding properties. Solid State Sci 2011;13:5538.
[28] Kobayashi Y, Shirochi T, Yasuda Y, Morita T. Metalmetal
bonding process using metallic copper nanoparticles
prepared in aqueous solution. Int J Adhes Adhes
2012;33:505.
[29] Kobayashi Y, Ishida S, Ihara K, Yasuda Y, Morita T.
Preparation of copper nanoparticles and metalmetal
bonding process using them. World J Eng 2013;10:
1138.
[30] Kobayashi Y, Shirochi T, Maeda T, Yasuda Y, Morita T.
Microstructure of metallic copper nanoparticles/metallic
disc interface in metalmetal bonding using them. Surf Int
Anal 2013;45:14248.

121

[31] Maeda T, Abe Y, Kobayashi Y, Yasuda Y, Morita T. Synthesis


of metallic copper nanoparticles using copper oxide
nanoparticles as precursor and their metalmetal bonding
properties. Sci Technol Weld Join 2012;17:48994.
[32] Kobayashi Y, Maeda T, Watanabe K, Ihara K, Yasuda Y, Morita
T. Preparation of CuO nanoparticles by metal saltbase
reaction in aqueous solution and their metallic bonding
property. J Nanoparticle Res 2011;13:536572.
[33] Morita T, Ide E, Yasuda Y, Hirose A, Kobayashi K. Study of
bonding technology using silver nanoparticles. Jpn J Appl
Phys 2008;47:661522.
[34] Morita T, Yasuda Y, Ide E, Akada Y, Hirose A. Bonding
technique using micro-scaled silver-oxide particles for
in-situ formation of silver nanoparticles. Mater Trans
2008;49:287580.
[35] Morita T, Yasuda Y, Ide E, Hirose A. Direct bonding to
aluminum with silver-oxide microparticles. Mater Trans
2009;50:2268.

You might also like