Next Article in Journal
Development of an Experimental Setup to Investigate Influences on Component Distortion in Gravity Die Casting and a First Variation of Temperature Control Strategy
Next Article in Special Issue
Inconel 713C Coating by Cold Spray for Surface Enhancement of Inconel 718
Previous Article in Journal
Fatigue Behavior of Al 7075-T6 Plates Repaired with Composite Patch under the Effect of Overload
Previous Article in Special Issue
Hardening Efficiency and Microstructural Changes during Laser Surface Hardening of 50CrMo4 Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Isothermal Aging on Interfacial Microstructure and Shear Properties of Sn-4.5Sb-3.5Bi-0.1Ag Soldering with ENIG and ENEPIG Substrates

1
Hebei Key Laboratory of Material Near-Net Forming Technology, School of Materials Science and Engineering, Hebei University of Science and Technology, Shijiazhuang 050018, China
2
China Electronics Technology Group Corp 13th Research Institute, Shijiazhuang 050061, China
*
Author to whom correspondence should be addressed.
Metals 2021, 11(12), 2027; https://doi.org/10.3390/met11122027
Submission received: 15 November 2021 / Revised: 8 December 2021 / Accepted: 12 December 2021 / Published: 14 December 2021
(This article belongs to the Special Issue Advances in Welding, Joining and Surface Coating Technology)

Abstract

:
Sn–Sb system solders and ENIG/ENEPIG surface finish layers are commonly used in electronic products. To illustrate the thermal reliability evaluation of such solder joints, we studied the interfacial microstructure and shear properties of Sn-4.5Sb-3.5Bi-0.1Ag/ENIG and Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG solder joints after aging at 150 °C for 250, 500 and 1000 h. The results show that the intermetallic compound of Sn-4.5Sb-3.5Bi-0.1Ag/ENIG interface was more continuous and uniform compared with that of Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG interface after reflow. The thickness of the interfacial intermetallic compounds of the former was significantly thinner than that of the latter before and after aging. With extension of aging time, the former interface was stable, while obvious voids appeared at the interface of the latter after 500 h aging and significant fracture occurred after 1000 h aging. The shear tests proved that shear strength of solder joints decreased with increasing aging time. For the Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG joint after 1000 h aging, the fracture mode is ductile-brittle mixed type, which means fracture could occur at the solder matrix or the solder/IMC interface. For other samples of these two types of joints, ductile fracture occurred inside of the solder. The Sn-4.5Sb-3.5Bi-0.1Ag/ENIG solder joint was thermally more reliable than Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG.

1. Introduction

Sn–Pb solder, as a connection material for electronic products, has always been favored by the electronics industry due to its excellent technological performance and low cost. However, because of the toxicity of Pb, countries in succession set off a wave of “lead-free” after entering the 21st century. At present, electronic manufacturing is in a special stage of transition from lead-containing to lead-free soldering. Among the several lead-free solders, the ternary alloy Sn–Ag–Cu is regarded as the leading candidate system for reflow soldering. However, the lower melt point of the alloy limits its application in high temperature. The Sn–Sb system solders perform better in terms of cost and high temperature applicability [1].
Due to the solid solution strengthening of Sb element and the formation of Sn–Sb intermetallic compound (IMC), Sn–Sb solders have good creep resistance and higher strength [2,3,4,5]. In order to further improve performance, ternary or quaternary alloys are developed by adding alloying elements. For example, the pronounced grain refinement is achieved as the addition of Ag [6]. Adding element of Bi led to solid solution strengthening and enhanced thermal reliability of Sn–Sb alloy [7].
Study of interfacial reactions between solders and metallization layers is a crucial subject for the reliability evaluation of the solder joints. Studies have shown that the stability of the solder joint interface largely depends on the IMC generated by the metallurgical reaction between the solder and the substrate surface during the soldering process [8,9,10]. After thermal aging, the thickness of the intermetallic compound will increase, and its own brittleness and the Kirkendall voids caused by the unequal diffusion of atoms during the growth process will severely weaken the mechanical properties of the solder joints. Therefore, it is very necessary to study the interface microstructure and mechanical properties of solder joints after isothermal aging. Cu is commonly used as a metallization layer in the soldering process. As Cu substrate contacts with solder directly during high temperature reflow soldering, the Cu element quickly dissolve and diffuse into the solder, which is prone to overreaction and affects the reliability of solder joints [11,12,13,14]. Thus surface treatments are applied to improve the interfacial properties of solder joints. Among many surface treatments, electroless nickel-immersion gold (ENIG) is the most mature one [15,16]. A layer of Ni is coated on the Cu substrate, and then a layer of Au is coated on the Ni layer, where Ni and Au coatings act as the barrier and anti-oxidation layer, respectively. Recently, the interfacial reaction between Sn—A–-Cu solder and ENIG surface finish has been widely studied. The results indicated that (Ni, Cu)3Sn4 phase formed at ENIG/Sn-3.5Ag-0.7Cu/ENIG sandwiched interfaces after reflowing. After aging at 150 °C for 24 h, the IMC became (Cu, Ni)6Sn5 phase. The interface composition did not further change after aging for 150 h. The IMC changed back to (Ni, Cu)3Sn4 phase after aging for 500 h [17]. The (Ni, Cu)3Sn4 and (Cu, Ni)6Sn5 interface compound were also found at Sn-4.0 wt % Ag-0.5 wt % Cu/ENIG interface, and voids were found inside the Sn–Ni–Cu ternary interfacial compounds after 500 and 1000 h aging [18].
However, due to the complexity of ENIG process, the phenomenon of “black pad” will appear in the process of actual use, which affects the reliability of soldering [19]. In addition to ENIG, electroless nickel-electroless palladium-immersion gold (ENEPIG) as a feasible surface finish has been widely used in various fields [20]. The Pd layer used as a protective layer to prevent the corrosion of the nickel (phosphorus) layer during the gold plating process and effectively solve the black pad problems [21]. Simultaneously, the interfacial reactions and mechanical properties of Solders/ENEPIG are extensively investigated. Chien-Fu et al. [22] studied the interfacial reaction and mechanical strength of Sn-3.0 wt % Ag-0.5 wt % Cu jointed with ENEPIG surface finish. The needle-like (Cu, Ni, Pd)6Sn5 formed in the ENEPIG joints. By inspecting the fracture, the ductile fracture was observed in the ENEPIG joint.
Nevertheless, data and reports on comparative studies between ENIG and ENEPIG for the interfacial reactions are still seriously lacking, especially combined with high-temperature Sn–Sb solder. In this work, the effects of isothermal aging on interfacial microstructure and shear properties of Sn-4.5Sb-3.5Bi-0.1Ag jointed with ENIG and ENEPIG surface finishes were investigated. The results would provide theoretical basis for evaluating the reliability of two solder joints.

2. Materials and Methods

In order to prepare the ENIG and ENEPIG substrates, the Au/Ni bilayer with the thickness of 0.1 μm/4 μm and the Au/Pd/Ni trilayer with the thickness of 0.1 μm/0.2 μm/4 μm were deposited on the Cu pads. The component is chip ceramic capacitor 0402, and the substrate is a 28 mm × 28 mm, 1.6 mm thick multi-layered (ground copper planes) FR-4 printed circuit board. The Sn-4.5Sb-3.5Bi-0.1Ag solder was printed on the ENIG or ENEPIG substrate by using an automatic solder paste and reflowed at a peak temperature of 280 °C for a belt speed of 1m/min in a vacuum reflow machine (Vacm-0023, Germany) to complete the surface mounting process. Here, the prepared Sn-4.5Sb-3.5Bi-0.1Ag/ENIG and Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG joints are named S1 and S2 respectively. The temperature profile of the reflow treatment and morphology of S1 after soldering was shown in Figure 1. In order to study their thermal aging properties, the two kinds of samples were aged in a high temperature test chamber (HT305E) at 150 °C for 250, 500 and 1000 h respectively.
The interfacial morphology of the joints was observed by scanning electron microscope (SEM; ZEISS ULTRA55, Carl Zeiss, Jena, Germany) after spraying gold on the surface of the specimens. To reveal the cross-sectional microstructure, the specimens were prepared by grounding, polishing and etching using 93%C2H5OH-5%HNO3-2%HCl etchant. Energy dispersive X-ray spectroscopy (EDS; Carl Zeiss, Jena, Germany) and micro-region X-ray diffraction (XRD; Bruker, Spabrucken, Germany) analysis were performed to identify the phases. The thickness of the IMC was carried out using image processing software (Adobe Photoshop CS6, Adobe, San Jose, America). The area pixel value A and length pixel value L of the intermetallic layer are measured by using Photoshop software, the formula h = A / L to find the number of pixel in the thickness of the intermetallic layer, and then obtain the actual length represented by each pixel according to the ratio of the scale value to the scale pixel value. Finally, the actual thickness of the interface IMC layer is obtained. The shear strength of solder joint was tested by MFM series bond tester, the shear height is 50 μm, and the shear speed is 700 μm/s. After shear testing, fracture and cross section morphologies were observed and analyzed by SEM and EDS.

3. Results and Discussion

3.1. Original Morphology and Compositions of IMC

Figure 2 shows cross-sectional SEM images and EDS results of the interfaces of S1 and S2. The EDS point analysis indicated that Ni3Sn4 layer formed at the interface of S1, which was continuous and uniform, showing a typical scallop shape, as presented in Figure 2a,b. At the interface of S2, a discontinuous but protruding massive IMC is formed. The EDS analysis indicated the massive IMC was in the (Pd, Ni)Sn4 phase, whose average atomic fractions are 78.14 at.%, Sn-4.38 at.%, and Ni-17.48 at.% Pd, as illustrated in Figure 2c,d. In addition, the detachment of the (Pd, Ni)Sn4 phase away from the Ni layer was found in Figure 2c.
For the two solder joints, no Au compounds were detected in the interface. Since the thickness of the Au coating layer is approximate 0.1 μm. At the beginning of the reflow, the Au layer quickly dissolves into the solder within a few seconds. Thus, the Ni layer of ENIG coating exposed to the solder side and reacted with molten solder to form the Ni3Sn4 phase. After Au atoms consumed out at the interface, Pd layer of the ENEPIG coating is sequentially exposed to the solder. Unlike Au surface layer, the thickness of the intermediate Pd layer is about twice as much as the Au layer, and the dissolution rate of Pd in the solder is about two orders of magnitude lower than the dissolution rate of Au [23]. The Pd layer reacted with Sn to form PdSn4 phase, as reported by Tanaka S et al. [24] There are extensive interaction between Pd and Ni, and the Ni atoms diffuse to the Pd-containing IMCs with a large flux. We speculate that as Pd atoms leaves the interface, Ni atoms can dissolve into the PdSn4 phase and replace the position of Pd atom to form (Pd, Ni)Sn4 phase.

3.2. The Impact of Aging Treatment on IMC

Figure 3 is cross-sectional SEM micrographs of the solder joints aged at 150 °C, where the left side shows interfacial microstructure of S1 after thermal aging. When the samples were subjected to thermal aging, they exhibited different microstructural evolution at the interfaces. The scallop shape morphology of the interface is no longer so obvious, and the entire interface tends to be flat after 250 and 500 h aging, illustrated by Figure 3a,b. Figure 4a shows the XRD pattern of the interface of S1 after aging for 250 h, the interface compound is still Ni3Sn4. When the aging time was extended to 1000 h, a new phase was formed at the interface, as seen by the arrows in Figure 3c. The EDS analysis indicated the average composition of this new phase was 7.34 at.% Cu-35.77 at.% Ni-56.89 at.% Sn, as presented in Figure 5, which was identified as the (Ni, Cu)3Sn4 phase. To explain how the Ni3Sn4 phase changed into (Ni, Cu)3Sn4 phase, we conducted series of EDS mapping analysis. Figure 6 is EDS mapping result of the interface of S1 after aging for 1000 h at 150 °C. It can be seen that the Ni layer becomes discontinuous, resulting in several gaps. We speculate that these gaps provide channels for diffusion of Cu atoms in the substrate and Cu can interact with Ni atoms at the interface. As Cu had similar crystal structure with Ni, Ni atoms were gradually replaced by the Cu atoms to form the (Ni, Cu)3Sn4 phase. The aging test proved that the interface of S1 is stable, and no voids or cracks were found.
The right side of Figure 3 shows the morphology of the interface of S2 after aging. Figure 4b is XRD pattern of the interface of S2 aged at 150 °C for 250 h, which indicates that Ni3Sn4 phase formed. Related research found that the (Pd, Ni)Sn4 phase will spall into molten solder after reflow, and as the reaction progresses, the spalled (Pd, Ni)Sn4 phase will dissolve into the molten solder [25]. As described earlier, the spalled massive (Pd, Ni)Sn4 phase was observed at the interface of S2, as seen in Figure 2c. But after aging, (Pd, Ni)Sn4 phase had completely dissolved into the solder. Simultaneously, the Ni layer exposed and Ni3Sn4 phase formed. Kirkendall voids are generated in IMC when aging at 150 °C for 500 h. The voids are so tiny that will only be discovered after long time aging. The cause of Kirkendall voids is due to the difference in the diffusion rate between atoms. This unbalanced diffusion mechanism leads to the generation of atomic-level holes, which is accompanied by the Kirkendall effect during the reflow process. The aging experiment accelerated the occurrence of the Kirkendall effect and prompted the Kirkendall voids to gather and grow, forming larger voids which are easy to be observed. A continuous fracture dramatically occurred along the interface after aging for 1000 h. The thermal stress is suggested to be affected by the cooling process, and during which the fracture initiated from the local voids and propagated across the entire interface.
The Figure 3d,h show the morphology of the Ni3P crystallization layer of S1 and S2 after aging 1000 h, respectively. The Ni3P crystallization layer is stable and no voids were found. According to the study of Tseng et al. [26], on the case of interfacial reaction of Sn–3.0Ag–0.5Cu solder jointed with ENIG and ENEPIG surface finishes, Ni3Sn4 was inhibited in the ENEPIG joints, which suppressed the excessive voids formation as compared to ENIG during thermal aging. In this case, Cu6Sn5 phase would form in the interfacial reaction of Sn–3.0Ag–0.5Cu solder with electroless Ni–P/immersion Au (ENIG) and electroless Ni–P/electroless Pd/immersion Au (ENEPIG). Moreover, Ni3Sn4 phase could occur in the ENIG aged joints, but not in the ENEPIG aged joints. It was demonstrated that the Ni3Sn4 phase usually accelerate the growth of columnar Kirkendall voids inside the Ni3P layer, thus resulting in poor performance of welded joints. However, the Sn–Sb–Ag solder used in this paper does not contain the Cu element, the Cu6Sn5 phase would not form at the interface of S1 and S2 and only the Ni3Sn4 phase formed in both aged joints. In addition to Ni3Sn4 phase, it is proved that the ratio of phosphorus to nickel in the Ni coating layer and the soldering procedure will also affect the formation of Ni3P layer. For Sn–Sb–Ag solder, we adopted reasonable soldering process that make Ni3P crystallization layer without columnar voids.
The IMC thickness of the solder joints increases with increasing aging time. It is generally believed that the rate controlling mechanism for the growth of the IMC is a diffusion process [27]. Figure 7 shows a function of the IMC thickness and the square root of the aging time. The IMC layer thickness increased linearly with the square root of aging time, which can be described by the following formula: X X 0 = D t , where the symbols X , X 0 , D and t represent the thickness of IMC layer after and before growth, growth rate constant and aging time, respectively. By calculation, the growth rate constants D of S1 and S2 interfacial compounds are 1.298 × 10−14 and 5.516 × 10−14 cm2/s, respectively. With increasing aging time, Ni atoms in ENIG and ENEPIG continuously diffuse into the solder, resulting in growth of the interface. Since the presence of Pd can increase the growth of IMC [28], thus the IMC thickness of S2 is thicker than that of S1.

3.3. Mechanical Properties

The relationship between shear strength and aging time is shown in Figure 8. The shear strength of the solder joint tends to be stable. After 1000 h aging, the shear strength of the S1 and S2 decreased by 5.48% and 10.85%, respectively. S2 exhibits lower shear strength than S1 during the aging process.
As we know, shear failure can occur in solder, solder/IMC interface or IMC, corresponding to ductile fracture, ductile-brittle mixed fracture and brittle fracture modes, respectively. Figure 9 shows the fracture morphologies of solder joints before and after different time aging at 150 °C. Figure 9a, f are the macroscopic fracture morphology of S1 and S2, respectively. The macroscopic morphology of the fracture is mainly composed of two parts. The fracture morphology at position 1 is basically the same, so we mainly analyzed the microscopic morphology of the fracture at position 2. The left-hand side columns show the fracture morphologies of S1. There are densely and small-sized dimple structures distributed at the fracture before aging, as indicated in Figure 9b. The dimples of the shear fracture became larger and the number decreased after aging for 250 h, as shown in Figure 9c. The dimples become lesser and local smooth areas appeared after aging for 500 h. And after 1000 h of aging, the local smooth areas have increased with increasing aging time.
Figure 10 shows the cross-sectional view of fracture surfaces of solder joints after different aging time under 150 °C. The left-hand side columns show the interfaces of S1. As presented in Figure 10a–d, the fracture position of the solder joint occurred in the solder matrix and gradually approaches the IMC interface. This trend was obvious in most tested samples. For each kind of solder joints aging at different time, five samples were analyzed. Due to the influence of artificial experimental operations, the five samples of each combination of solder alloy have different fracture positions in the solder matrix, but with the increase of aging time, the fracture position of the samples show a tendency to gradually approach the IMC layer. Related studies indicate that for the aged samples, as the aging time increases, the shear fracture of the solder joint usually occurs in the IMC layer [29]. However, Figure 10d shows that the shear fracture of the S1 still occurred in the solder after aging for 1000 h. On the one hand, IMC thickness of this solder joint is relatively thin, and the morphology is flat, resulting in a more uniform stress distribution. On the other hand, Cu–Ni–Sn IMCs have high shear modulus [30]. Therefore, the shear forces tears the solder matrix of S1 into elongated dimples, and the failure mode is still ductile fracture after long time aging.
The right-hand side columns show the fracture morphologies of S2 before and after different time aging at 150 °C. Same to the fracture morphology of S1, there is densely and small-sized dimple structures distributed at the fracture surface of S2 before aging, as indicated in Figure 9g. However, local smooth areas appeared after aging for 250 h and there are nearly no dimples on the fracture surface after 500 h aging. Moreover, it can be found from Figure 9j that pits were formed on the fracture surface of the solder joint which was aged for 1000 h, and the bottom of pits were composed of Ni3Sn4 phase. Combining with Figure 10h, it can conclude that one part of the fracture occurred at the interface, and the other part occurred in the solder matrix. The fracture mode of the solder joint is ductile-brittle mixed, with the lowest strength. The IMC thickness of S2 is higher than S1 before and after aging, and the IMC shape of the former is irregular. The thicker IMC layer and irregular shape of the IMC increase the roughness of the S2 interface, which causes large local stress of the interface under the shear force and partial fracture on the solder/IMC interface.

4. Conclusions

In this study, a comparative analysis of interfacial microstructure and shear properties of Sn-4.5Sb-3.5Bi-0.1Ag solder with two kinds of surface finish layers (ENIG and ENEPIG) was conducted. The conclusions are as follows:
(1) For the Sn-4.5Sb-3.5Bi-0.1Ag/ENIG interface, a continuous and uniform scallop-like Ni3Sn4 interface compound is formed after reflow. With the increase of aging time, the thickness of the interface IMC layer gradually increased, and no obvious voids and fracture generated at interface. When the aging time was 1000 h, the initial thickness increased from 1.86 μm to 4.05 μm, and the interface is consisted of (Ni, Cu)3Sn4 phase.
(2) For the Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG interface, a massive discontinuous (Pd, Ni)Sn4 phase formed after reflow. After aging, the IMC consisted of the Ni3Sn4 phase. With the increasing of aging time, the interface thickness increases continuously and Kirkendall voids are generated. After aging for 1000 h, a continuous fracture occurred.
(3) The shear strength of Sn-4.5Sb-3.5Bi-0.1Ag/ENIG and Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG solder joints decrease with the increase of thermal aging time. After 1000 h aging, the shear strength of the two solder joints decreased by 5.48% and 10.85%, respectively. The fracture mode of the Sn-4.5Sb-3.5Bi-0.1Ag/ENIG solder joint is ductile after reflow and aging. The fracture type of Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG solder joint is ductile after reflow and 250, 500 h aging. When the aging time was extended to 1000 h, the fracture pattern transformed into ductile-brittle mixed mode.
As stated previously, the Sn-4.5Sb-3.5Bi-0.1Ag/ENIG solder joint is thermally more reliable than Sn-4.5Sb-3.5Bi-0.1Ag/ENEPIG.

Author Contributions

Conceptualization, Z.L. and Z.P.; methodology, Z.L., Z.P. and L.W.; validation, Z.L., Z.P., F.S. and D.W.; investigation, F.S., Z.Y., D.X. and S.W.; writing—original draft preparation, Z.L., Z.P., F.S. and Z.Y.; writing—review and editing, Z.P., Z.L. and F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Key Research and Development Program of Hebei Province (Grant No. 19250202D), National Natural Science Foundation of China (Grant No. 51875168), Natural Science Foundation of Hebei Province (E2020208083).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request.

Acknowledgments

The authors would like to thank for the support and facilitation, Hebei Iron and Steel Technology Research Institute.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ENIGElectroless Nickel-Immersion Gold
ENEPIGElectroless Nickel-Electroless Palladium-Immersion Gold
Sn Tin
SbAntimony
BiBismuth
AgSilver
CuCopper
NiNickel
Pd Palladium
Au Gold
SEMScanning Electron Microscopy
EDSEnergy Dispersive X-Ray Spectroscopy
XRDX-ray diffraction
IMCIntermetallic compound

References

  1. Puttlitz, K.J.; Stalter, K.A. Handbook of Lead-Free Solder Technology for Microelectronic Assemblies; Marcel Dekker, Inc.: New York, NY, USA, 2004. [Google Scholar]
  2. Murty, K.L.; Haggag, F.M.; Mahidhara, R.K. Tensile, creep, and ABI tests on sn5%sb solder for mechanical property evaluation. J. Mater. 1997, 26, 839–846. [Google Scholar] [CrossRef]
  3. McCabe, R.J.; Fine, M.E. Creep of tin, Sb-solution-strengthened tin, and SbSn-precipitate-strengthened tin. Metall. Mater. A 2002, 33, 1531–1539. [Google Scholar] [CrossRef]
  4. Mahmudi, R.; Geranmayeh, A.R.; Bakherad, M.; Allami, M. Indentation creep study of lead-free Sn-5%Sb solder alloy. Mater. Sci. Eng. A 2007, 457, 173–179. [Google Scholar] [CrossRef]
  5. Mahmudi, R.; Geranmayeh, A.R.; Rezaee-Bazzaz, A. Impression creep behavior of lead-free Sn-5Sb solder alloy. Mater. Sci. Eng. A 2007, 448, 287–293. [Google Scholar] [CrossRef]
  6. El-Daly, A.A.; Fawzy, A.; Mohamad, A.Z.; El-Taher, A.M. Microstructural evolution and tensile properties of Sn-5Sb solder alloy containing small amount of Ag and Cu. J. Alloys Compd. 2011, 509, 4574–4582. [Google Scholar] [CrossRef]
  7. Geranmayeh, A.R.; Nayyeri, G.; Mahmudi, R. Microstructure and impression creep behavior of lead-free Sn-5Sb solder alloy containing Bi and Ag. Mater. Sci. Eng. A 2012, 547, 110–119. [Google Scholar] [CrossRef]
  8. Duan, N.; Scheer, J.; Bielen, J.; van Kleef, M. The influence of Sn-Cu-Ni(Au) and Sn-Au intermetallic compounds on the solder joint reliability of flip chips on low temperature co-fired ceramic substrates. Microelectron. Reliab. 2003, 43, 1317–1327. [Google Scholar] [CrossRef]
  9. Zeng, K.; Tu, K.N. Six cases of reliability study of Pb-free solder joints in electronic packaging technology. Mater. Sci. Eng. R 2002, 38, 55–105. [Google Scholar] [CrossRef]
  10. Foley, J.C.; Gickler, A.; Leprevost, F.H.; Brown, D. Analysis of ring and plug shear strengths for comparison of lead-free solders. J. Mater. 2000, 29, 1258–1263. [Google Scholar] [CrossRef]
  11. Gong, J.C.; Liu, C.Q.; Conway, P.P.; Silberschmidt, V.V. Evolution of CuSn intermetallics between molten SnAgCu solder and Cu substrate. Acta. Mater. 2008, 56, 4291–4297. [Google Scholar] [CrossRef]
  12. Wang, F.J.; Yu, Z.S.; Qi, K. Intermetallic compound formation at Sn-3.0Ag-0.5Cu-1.0Zn lead-free solder alloy/Cu interface during as-soldered and as-aged conditions. J. Alloy. Compd. 2007, 438, 110–115. [Google Scholar] [CrossRef]
  13. Lee, Y.H.; Lee, H.T. Shear strength and interfacial microstructure of Sn-Ag-xNi/Cu single shear lap solder joints. Mater. Sci. Eng. A 2007, 444, 75–83. [Google Scholar] [CrossRef]
  14. Zhang, R.; Tian, Y.H.; Hang, C.J.; Liu, B.L.; Wang, C.Q. Formation mechanism and orientation of Cu3Sn grains in Cu-Sn intermetallic compound joints. Mater. Lett. 2013, 110, 137–140. [Google Scholar] [CrossRef]
  15. Sohn, Y.C.; Yu, J. Correlation between chemical reaction and brittle fracture found in electroless Ni(P)/immersion gold–solder interconnection. J. Mater. Res. 2005, 20, 1931–1934. [Google Scholar] [CrossRef]
  16. Yoon, J.W.; Noh, B.I.; Jung, S.B. Comparison of Interfacial Stability of Pb-Free Solders (Sn-3.5Ag, Sn-3.5Ag-0.7Cu, and Sn-0.7Cu) on ENIG-Plated Cu During Aging. IEEE. Trans. Compon. Packag. Technol. 2010, 33, 64–70. [Google Scholar] [CrossRef]
  17. Yoon, J.W.; Moon, W.C.; Jung, S.B. Interfacial reaction of ENIG/Sn-Ag-Cu/ENIG sandwich solder joint during isothermal aging. Microelectron. Eng. 2006, 83, 2329–2334. [Google Scholar] [CrossRef]
  18. Sun, P.; Andersson, C.; Wei, X.C.; Cheng, Z.N.; Shangguan, D.K.; Liu, J. High temperature aging study of intermetallic compound formation of Sn-3.5Ag and Sn-4.0Ag-0.5Cu solders on electroless Ni(P) metallization. J. Alloys Compd. 2006, 425, 191–199. [Google Scholar] [CrossRef]
  19. Snugovsky, P.; Arrowsmith, P.; Romansky, M. Electroless Ni/Immersion Au interconnects: Investigation of black pad in wire bonds and solder joints. J. Electron. Mater. 2001, 30, 1262–1270. [Google Scholar] [CrossRef]
  20. Yoon, J.W.; Noh, B.I.; Yoon, J.H.; Kang, H.B.; Jung, S.B. Sequential interfacial intermetallic compound formation of Cu6Sn5 and Ni3Sn4 between Sn-Ag-Cu solder and ENEPIG substrate during a reflow process. J. Alloys Compd. 2011, 509, 153–156. [Google Scholar] [CrossRef]
  21. Kim, P.G.; Tu, K.N.; Abbott, D.C. Time-and temperature-dependent wetting behavior of eutectic SnPb on Cu leadframes plated with Pd/Ni and Au/Pd/Ni thin films. J. Appl. Phys. 1998, 84, 770–775. [Google Scholar] [CrossRef]
  22. Tseng, C.F.; Duh, J.G. Correlation between microstructure evolution and mechanical strength in the Sn-3.0Ag-0.5Cu/ENEPIG solder joint. Mater. Sci. Eng. A 2013, 580, 169–174. [Google Scholar] [CrossRef]
  23. Bader, W.G. Dissolution of Au, Pd, Pt, Cu and Ni in a molten tin-lead solder. Weld. Res. Supp. 1969, 48, 551–557. [Google Scholar]
  24. Tanaka, S.; Kajihara, M. Kinetics of solid-state reactive diffusion between Sn-Ni alloys and Pd. J. Alloys Compd. 2009, 484, 273–279. [Google Scholar] [CrossRef]
  25. Baek, Y.H.; Chung, B.M.; Choi, Y.S.; Choi, J.H.; Huh, J.Y. Effects of Ni3Sn4 and (Cu,Ni)6Sn5 intermetallic layers on cross-interaction between Pd and Ni in solder joints. J. Alloys Compd. 2013, 579, 75–81. [Google Scholar] [CrossRef]
  26. Tseng, C.F.; Lee, T.K.; Ramakrishna, G.; Liu, K.C.; Duh, J.G. Suppressing Ni3Sn4 formation in the Sn–Ag–Cu solder joints with Ni–P/Pd/Au surface finish. Mater. Lett. 2011, 65, 3216–3218. [Google Scholar] [CrossRef]
  27. Yoon, J.W.; Lee, C.B.; Jung, S.B. Growth of an intermetallic compound layer with Sn-3.5Ag-5Bi on Cu and Ni-P/Cu during aging treatment. J. Electron. Mater. 2003, 32, 1195–1202. [Google Scholar] [CrossRef]
  28. Cao, L.H.; Chen, Y.B.; Shi, Y.Y. Effects of Alloy Elements on the Interfacial Microstructure and Shear Strength of Sn-Ag-Cu Solder. Acta Metall Sin 2019, 55, 1606–1614. [Google Scholar]
  29. Ahat, S.; Mei, S.; Le, L. Microstructure and shear strength evolution of SnAg/Cu surface mount solder joint during aging. J. Electron. Mater. 2001, 30, 1317–1322. [Google Scholar] [CrossRef]
  30. Che, F.X.; Pang, J. Characterization of IMC layer and its effect on thermomechanical fatigue life of Sn-38Ag-0.7Cu solder joints. J. Alloys Compd. 2012, 541, 6–13. [Google Scholar] [CrossRef]
Figure 1. Temperature profile of the reflow treatment (a) and the scanning electron microscope (SEM) image shows morphology of S1 (b).
Figure 1. Temperature profile of the reflow treatment (a) and the scanning electron microscope (SEM) image shows morphology of S1 (b).
Metals 11 02027 g001
Figure 2. Cross-sectional SEM images and EDS results of S1 (a,b) and S2 (c,d).
Figure 2. Cross-sectional SEM images and EDS results of S1 (a,b) and S2 (c,d).
Metals 11 02027 g002
Figure 3. Cross-sectional SEM micrographs of the solder joints aged at 150 °C. (ac) The interface morphologies of S1 after aging for 250, 500 and 1000 h, respectively. (d) Dense Ni3P layer on the interface of S1. (eg) The interface morphologies of S2 after aging for 250, 500 and 1000 h, respectively. (h) Dense Ni3P layer on the interface of S2.
Figure 3. Cross-sectional SEM micrographs of the solder joints aged at 150 °C. (ac) The interface morphologies of S1 after aging for 250, 500 and 1000 h, respectively. (d) Dense Ni3P layer on the interface of S1. (eg) The interface morphologies of S2 after aging for 250, 500 and 1000 h, respectively. (h) Dense Ni3P layer on the interface of S2.
Metals 11 02027 g003
Figure 4. XRD patterns of the interfaces of (a) S1 and (b) S2 aged at 150 °C for 250 h.
Figure 4. XRD patterns of the interfaces of (a) S1 and (b) S2 aged at 150 °C for 250 h.
Metals 11 02027 g004
Figure 5. EDS pattern of the interface of S1 aged for 1000 h.
Figure 5. EDS pattern of the interface of S1 aged for 1000 h.
Metals 11 02027 g005
Figure 6. EDS mapping analysis of the interface of S1 aged at 150 °C for 1000 h.
Figure 6. EDS mapping analysis of the interface of S1 aged at 150 °C for 1000 h.
Metals 11 02027 g006
Figure 7. IMC thickness as a function of aging time.
Figure 7. IMC thickness as a function of aging time.
Metals 11 02027 g007
Figure 8. Shear strength of solder joints after different aging time under 150 °C.
Figure 8. Shear strength of solder joints after different aging time under 150 °C.
Metals 11 02027 g008
Figure 9. The left-hand and right-hand side columns show the fracture morphologies of S1 and S2 after aging different time at 150 °C. (a) Macroscopic fracture image of S1. (be) Microscopic morphologies of fractured S1 after reflow and after aging for 250, 500, 1000 h, respectively. (f) Macroscopic fracture image of S2. (gj) Microscopic morphologies of fractured S2 after reflow and after aging for 250, 500, 1000 h, respectively.
Figure 9. The left-hand and right-hand side columns show the fracture morphologies of S1 and S2 after aging different time at 150 °C. (a) Macroscopic fracture image of S1. (be) Microscopic morphologies of fractured S1 after reflow and after aging for 250, 500, 1000 h, respectively. (f) Macroscopic fracture image of S2. (gj) Microscopic morphologies of fractured S2 after reflow and after aging for 250, 500, 1000 h, respectively.
Metals 11 02027 g009
Figure 10. The left-hand and right-hand side columns show the cross-sectional view of fractured S1 and S2 (ad) Morphologies of fractured S1 after reflow and after aging for 250, 500, 1000 h under 150 °C, respectively. (eh) Morphologies of fractured S2 after reflow and after aging for 250, 500, 1000 h under 150 °C, respectively.
Figure 10. The left-hand and right-hand side columns show the cross-sectional view of fractured S1 and S2 (ad) Morphologies of fractured S1 after reflow and after aging for 250, 500, 1000 h under 150 °C, respectively. (eh) Morphologies of fractured S2 after reflow and after aging for 250, 500, 1000 h under 150 °C, respectively.
Metals 11 02027 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liang, Z.; Shen, F.; Yang, Z.; Xu, D.; Wei, S.; Peng, Z.; Wang, L.; Wang, D. Effects of Isothermal Aging on Interfacial Microstructure and Shear Properties of Sn-4.5Sb-3.5Bi-0.1Ag Soldering with ENIG and ENEPIG Substrates. Metals 2021, 11, 2027. https://doi.org/10.3390/met11122027

AMA Style

Liang Z, Shen F, Yang Z, Xu D, Wei S, Peng Z, Wang L, Wang D. Effects of Isothermal Aging on Interfacial Microstructure and Shear Properties of Sn-4.5Sb-3.5Bi-0.1Ag Soldering with ENIG and ENEPIG Substrates. Metals. 2021; 11(12):2027. https://doi.org/10.3390/met11122027

Chicago/Turabian Style

Liang, Zhimin, Fei Shen, Zongyuan Yang, Da Xu, Shaowei Wei, Zhenzhen Peng, Liwei Wang, and Dianlong Wang. 2021. "Effects of Isothermal Aging on Interfacial Microstructure and Shear Properties of Sn-4.5Sb-3.5Bi-0.1Ag Soldering with ENIG and ENEPIG Substrates" Metals 11, no. 12: 2027. https://doi.org/10.3390/met11122027

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop