2019 Volume 60 Issue 6 Pages 876-881
In this study, evaluation of fatigue crack properties in the Sn–5.0Sb/Cu joint was performed by using the inelastic strain energy density Win around the crack tip which was calculated by FEM. Win-c given by the multiplication of Win obtained from an arbitrary-sized square region surrounding the crack tip by the side length of the square region – did not depend on the size of the square region and the element size. The fatigue crack propagated in the solder layer and the power law of Paris law type between its fatigue crack propagation rate and ΔWin-c held. However, the power exponent in the fatigue crack propagation law differed depending on the regions of ΔWin-c. The power exponent became about 1 in the low ΔWin-c region and very large in the high ΔWin-c region. In the low ΔWin-c region, fatigue fracture propagated along the high angle grain boundaries formed ahead of the crack by the continuous dynamic recrystallization. On the other hand, the fracture transformed to static fracture mode in the high ΔWin-c region and the cleavage fracture was observed. The large power exponent in the high ΔWin-c region was attributed to the cleavage fracture.
Fatigue crack propagation rate as a function of ΔWin-c.
Power semiconductor modules using wide-band gap semiconductors, such as SiC are expected to operate at higher temperatures than in conventional conditions. Along with an increase in the operating temperature, high thermostability of materials that compose a power semiconductor module is now an important issue. Especially, the realization of high thermostability of die attach materials used for joining semiconductors, insulated substrates, and heat dissipating structure members is a matter of urgent attention. In the context, Sn–Sb based alloys are now drawing attention as a candidate for die-attach material. Die-attach materials are subjected to cyclic thermal stress ascribable to the differences in thermal expansion coefficients of members of the power module, which induces a fatigue crack to propagate, reducing heat dissipation characteristics and failure of the module is caused in the end. Therefore, evaluation of fatigue crack propagation rate in die attach materials is critical to guarantee reliability of power modules.
Since die-attach materials are used in the environment beyond small scale yielding, ideal for the evaluation of fatigue crack propagation rate is to use the cyclic J-integral ΔJ.1) Although fatigue crack propagation in solder alloys and Ag nanoparticles for the die-attach have been evaluated by ΔJ,2–4) the method of calculating ΔJ by the finite element method (FEM) is accompanied by difficulties and there is a problem to consider creep deformation into ΔJ. With these taken into consideration, evaluation by a scalar parameter as an alternate is desired. As a method without using a fracture mechanics parameter, there have been efforts to calculate the inelastic strain energy density range ΔWin around the crack tip by the FEM and the result has been used as a parameter to evaluate the fatigue crack propagation.5,6) As it is easy to calculate ΔWin by FEM, ΔWin can be expected as an alternate to ΔJ. However, since unlike the path-independent J-integral, ΔWin is dependent on the element size of FEM model and the distance from the crack tip, the dependence of ΔWin needs to be considered. Because of the foregoing, the author et al. have investigated a method to reduce the dependence of ΔWin on the element size and the distance from the crack tip.7)
In this study, fatigue crack propagation characteristic of Sn–5.0 mass%Sb that is candidate materials for die attach of power modules was evaluated by the method using inelastic strain energy density.
Sn–5.0 mass%Sb (hereafter mass% is omitted) was used in this study. A flat plate scarf joint with 300 µm of solder layer width as shown in Fig. 1 was employed for specimen of fatigue crack propagation testing. The joining angle was 45 degrees from the stress axis to produce the mixed mode loading of mode I and mode II. A notch was created on one side at the end of the joint. The notch with the length of 500 µm was introduced by a precision wire saw (K. D. Unipress). The joining temperature was +50 K above the melting temperature and soldering was done in the atmosphere with the use of RA flux. Then, the specimen surface was polished with SiC grinding papers and diamond particles. Finally, the surface was mirror-finished by polishing with colloidal silica suspension. Figure 2 shows the initial microstructure of the joint. The joint consists of few crystalline grains of β-Sn solid solution with Cu6Sn5 dispersoids. Although it is considered that Sb forms a solid solution in β-Sn and Sb exceeding the solubility limit forms SnSb intermetallic compounds, the compounds were not observed under the optical microscope.
Specimen geometry for fatigue crack propagation test.
Low magnification optical micrograph around the notch tip (left side) and high magnification optical micrograph in front of the notch tip (right side) of Sn–5.0Sb/Cu joint.
Fatigue crack propagation tests were performed in pulsating displacement-controlled tensile-tensile mode. The control waveform was symmetric triangle wave, the test temperatures were 348 K and 398 K, and the total displacement range was 5 µm–11 µm, and the displacement rate was 6 µm s−1. Figure 3 shows the testing system used in this study. The test machine shown in Fig. 3 is a micro-load fatigue test machine (Saginomiya Seisakusho: LMH207-20) equipped with a piezo actuator which was controlled by the displacements measured by the capacitance sensor fixed below the specimen. The test temperatures were controlled by ceramic heater installed inside the specimen fixing jigs. Fatigue crack lengths were measured from video images recorded by a microscope (Leica: Z16 APO) installed on the upper part of the test machine.
Fatigue crack propagation testing machine.
In fracture mechanics, the cyclic J-integral ΔJ shown in eq. (1) is used as the driving force for fatigue crack propagation in the condition beyond small scale yielding.1)
\begin{equation} \Delta J=\int_{\Gamma}\left(\varPsi(\Delta\varepsilon_{ij})dy-\Delta T_{i}\frac{\partial\Delta u_{i}}{\partial x}ds\right) \end{equation} | (1) |
\begin{equation} \dot{\chi}_{i}=\frac{2}{3}\sum_{i=1}^{n}A_{1}\dot{\varepsilon}_{\text{p}}-A_{2}\chi_{i}\dot{p} \end{equation} | (2) |
\begin{equation} \Delta W_{\text{in}}=\frac{\displaystyle\sum\Delta W_{\text{in}}^{\text{element}}\times V^{\text{element}}}{\displaystyle\sum V^{\text{element}}} \end{equation} | (3) |
FEM model for calculation of ΔWin.
Microstructures were observed with an optical microscope (Carl Zeiss: Axio Imager A1m). Furthermore, crystallographic orientation analysis was performed, using the technique of Electron Back Scatter Diffraction (EBSD, TSL: OIM). Specimens were prepared by mechanical polishing using SiC polishing paper and diamond paste, followed by further polishing using a colloidal silica suspension to remove the layers that were damaged in the mechanical polishing process.
Since ΔWin is dependent on the element size in the FEM model and the distance from the crack tip, these types of dependence need to be clarified. The HRR singularity field is formed surrounding the crack tip and the stress and the strain in the singularity field can be described by eq. (4) and eq. (5).11,12)
\begin{equation} \sigma_{ij}=k_{1}\left(\frac{J}{r}\right)^{\frac{1}{n+1}}{}\cdot\tilde{\sigma}_{ij}(\theta) \end{equation} | (4) |
\begin{equation} \varepsilon_{ij}=k_{2}\left(\frac{J}{r}\right)^{\frac{n}{n+1}}{}\cdot\tilde{\varepsilon}_{ij}(\theta) \end{equation} | (5) |
Illustration showing HRR singularity and Win acquisition area around crack tip in FEM analysis.
Figure 6 shows the relationship between Win and Larea for each Lelement. The gradient of a straight line became −1. As predicted above, Win was the power of −1 to Larea, i.e. becomes inversely proportional. As for this inversely proportional relationship, similar results had been obtained in sintered Ag nanoparticles7,13) and the inverse relationship can be established regardless of materials. Furthermore, Win calculated as an average in the square area has negligible effects of element size in the square area. Since Win and Larea are inversely proportional to each other, the proportional constant of a straight line is determined from Win calculated in an arbitrary size of square area as in eq. (6).
\begin{equation} W_{\text{in-c}}=W_{\text{in}}\cdot L_{\text{area}} \end{equation} | (6) |
\begin{align} C^{*}&=\int_{\Gamma}\left(W^{*}dy-\sigma_{ij}n_{j}\frac{\partial\dot{u}_{i}}{\partial x}ds\right)\\ W^{*}&=\int_{0}^{\dot{\varepsilon}_{kl}}\sigma_{ij}d\dot{\varepsilon}_{ij} \end{align} | (7) |
\begin{equation} \sigma_{ij}=\left(\frac{C^{*}}{AI_{n}r}\right)^{\frac{1}{n+1}}{}\cdot\tilde{\sigma}_{ij}(\theta) \end{equation} | (8) |
\begin{equation} \varepsilon_{ij}=\left(\frac{C^{*}}{AI_{n}r}\right)^{\frac{n}{n+1}}{}\cdot\tilde{\varepsilon}_{ij}(\theta) \end{equation} | (9) |
\begin{equation} \dot{\varepsilon}_{\text{ss}}=A_{3}\sigma^{n} \end{equation} | (10) |
Relationship between Win and Larea for each Lelement (elasto-plasticity analysis).
The relationship between Win and Larea is shown in Fig. 7. As in elasto-plastic analysis, Win has almost no effects of element size and becomes inversely proportional to Larea. Therefore, even in the condition in which creep deformation becomes dominant, fatigue crack propagation can be evaluated by using Win-c calculated by eq. (6).
Relationship between Win and Larea in for each Lelement (elasto-creep analysis).
In this study, fatigue crack propagation rate of Sn–5.0Sb was evaluated by Paris type fatigue crack propagation law using ΔWin-c shown in eq. (11).
\begin{equation} \frac{da}{dN}=C_{1}\Delta W_{\text{in-c}}{}^{C_{2}} \end{equation} | (11) |
The relationship between fatigue crack length and the number of cycles is shown in Fig. 8. It is clear that with an increase in the total displacement range, Win around the crack tip becomes large, causing the fatigue crack propagation rate to increase. The fatigue crack propagation rate da/dN was calculated from the fatigue crack propagation curve and the results organized by using ΔWin-c are shown in Fig. 9. There is variability among the test results, but a power law that should be shown in eq. (11) holds between the fatigue crack propagation rate and ΔWin-c. The power exponent of the fatigue crack propagation law differs depending on the levels of ΔWin-c and beyond the point exceeding ΔWin-c = 100 N m−1. The power exponent became about 1 in the low ΔWin-c region and more than 3 in the high ΔWin-c region. It has been reported that when fatigue crack propagation rate of solder alloys is evaluated by ΔJ, the power exponent of the fatigue crack propagation law becomes about 1.2) In addition, according to a research report on heat resistance alloys, the power exponent of the fatigue crack propagation law in the time-dependent type fatigue becomes 1.15) From these research reports, the exponent of fatigue crack propagation law in the time-dependent type fatigue becomes 1 regardless of the kind of material. As the test temperature of this study is more than 60% of the melting temperature of Sn–5.0Sb, the time-dependent type fatigue is dominant for the test conditions in this study. It is thought to be that the power exponent in the low ΔWin-c region showed the value of the time-dependent type fatigue.
Relationship between fatigue crack length and number of cycles for each testing condition.
Fatigue crack propagation rate as a function of ΔWin-c.
On the other hand, a large power exponent was obtained in the high ΔWin-c region and its value is very large as a ductile solder alloy. In the high ΔWin-c region, failure mode transcends the region of continuous crack propagation caused by fatigue and seems to become crack propagation caused by static fracture mode. And there are no effects of temperatures on the fatigue crack propagation rate.
3.2.2 Mechanism on fatigue crack propagationShown in Fig. 10 are bright-field images of fatigue crack propagation paths in the tests in the low and high regions of ΔWin-c. The fatigue crack propagated in the solder bulk in any ΔWin-c region. In the low ΔWin-c region, the fatigue crack deflected microscopically. In the high ΔWin-c region, the crack propagated linearly. The fracture mechanism varies according to the regions of ΔWin-c, which relates to the change in the exponent in the fatigue crack propagation law.
Optical micrographs showing crack propagation path in low ΔWin-c and high ΔWin-c region.
Figure 11 shows polarized light image around a fatigue crack in the low ΔWin-c region, and the crystalline orientation mapping and the grain boundary structure image by EBSD. In the grain boundary structure image, the red lines are subgrain boundaries with the misorientation less than 15 degrees and the blue lines are high angle grain boundaries with the misorientation more than 15 degrees. In the low ΔWin-c region, fine grains exist around the fatigue crack, and most of them were high angle grain boundaries. It has been reported that β-Sn, which is the matrix of Sn alloy has microstructural changes called “continuous dynamic recrystallization” during high temperature cyclic deformation.16–18) The continuous dynamic recrystallization is a phenomenon in which subgrains are formed in arranging dislocations through dynamic recovery and the continuous deformation increases the dislocation density, which makes the subgrains continuously rotate causing recrystallization to progress.19) In the condition of the low ΔWin-c region in this study, the continuous dynamic recrystallization occurred. Figure 12 shows the relationship between fatigue crack length and the number of cycles in the test of low ΔWin-c region. A staircase pattern was observed as shown in Fig. 12 and it is confirmed from the video records that in the phenomenon, crack propagation and non-propagation were repeated. The result shows that during the period of forming subgrains around the crack tip, propagation of the crack ceases and when a high angle grain boundary is formed by the rotating grains, the fatigue crack propagates along the high angle grain boundaries. In other words, it can be considered that the continuous dynamic recrystallization and subsequent grain boundary fracture in the low ΔWin-c region are the fatigue crack propagation mechanism.
Polarized light images, orientation mapping and grain boundary structure of around fatigue crack in low ΔWin-c region.
Relationship between fatigue crack length and number of cycles in low ΔWin-c condition.
Figure 13 shows polarized light image around the fatigue crack in the high ΔWin-c region, crystalline orientation mapping and grain boundary structure. In the images, the continuous dynamic recrystallization observed in the low ΔWin-c region was not seen in the high ΔWin-c region, but the cleavage fracture in grain interior was observed. Under a large strain energy condition, the fatigue crack propagation does not seem to be the continuum crack propagation mechanism by fatigue, but seems to become crack propagation by the static fracture mode mechanism. In this study, from ΔWin-c = 100 N m−1 onwards, the fracture mechanism seems to transit to the static fracture mode mechanism. The crack propagated in a brittle manner which was prompted by the cleavage fracture before the occurrence of the continuous dynamic recrystallization. The cleavage fracture contributes to increase the power exponent beyond the point exceeding ΔWin-c = 100 N m−1.
Polarized light images, orientation mapping and grain boundary structure around fatigue crack in high ΔWin-c region.
This paper is based on results obtained from a project commissioned by the New Energy and Industrial Technology Development Organization (NEDO).