Journal of Materials Science & Technology, 2020, 50(0): 115-127 DOI: 10.1016/j.jmst.2019.12.036

Research Article

A data-driven framework to predict the morphology of interfacial Cu6Sn5 IMC in SAC/Cu system during laser soldering

Anil Kunwar,a,b,*, Lili Anc, Jiahui Liuc, Shengyan Shangc, Peter Råback4, Haitao Mac, Xueguan Songb

aDepartment of Materials Engineering, KU Leuven, Kasteelpark Arenberg 44, B-3001 Leuven, Belgium

bSchool of Mechanical Engineering, Dalian University of Technology, Dalian 116024, China

cSchool of Materials Science and Engineering, Dalian University of Technology, Dalian 116024, China

dCSC-IT Center for Science, Keilaranta 14, P.O. Box 405, FIN-02101 Espoo, Finland

Corresponding authors: * Department of Materials Engineering, KU Leuven,Kasteelpark Arenberg 44, B-3001 Leuven, Belgium.E-mail address:kunwaranil@gmail.com(A. Kunwar).

Received: 2019-11-10   Accepted: 2019-12-18   Online: 2020-08-1

Abstract

A data-driven approach combining together the experimental laser soldering, finite element analysis and machine learning, has been utilized to predict the morphology of interfacial intermetallic compound (IMC) in Sn-xAg-yCu/Cu (SAC/Cu) system. Six types of SAC solders with varying weight proportion of Ag and Cu, have been processed with fiber laser at different magnitudes of power (30-50 W) and scan speed (10-240 mm/min), and the resultant IMC morphologies characterized through scanning electron microscope are categorized as prismatic and scalloped ones. For the different alloy composition and laser parameters, finite element method (FEM) is employed to compute the transient distribution of temperature at the interface of solder and substrates. The FEM-generated datasets are supplied to a neural network that predicts the IMC morphology through the quantified values of temperature dependent Jackson parameter (αJ). The numerical value of αJ predicted from neural network is validated with experimental IMC morphologies. The critical scan speed for the morphology transition between prismatic and scalloped IMC is estimated for each solder composition at a given power. Sn-0.7Cu having the largest critical scan speed at 30 W and Sn-3.5Ag alloy having the largest critical scan speed at input power values of 40 W and 50 W, thus possessing the greatest likelihood of forming prismatic interfacial IMC during laser soldering, can be inferred as most suitable SAC solders in applications exposed to shear loads.

Keywords: Intermetallic compound ; Neural network ; Finite element method (FEM) ; Laser parameters ; Lead-free solders ; Morphology

PDF (3503KB) Metadata Metrics Related articles Export EndNote| Ris| Bibtex  Favorite

Cite this article

Anil Kunwar, Lili An, Jiahui Liu, Shengyan Shang, Peter Råback, Haitao Ma, Xueguan Song. A data-driven framework to predict the morphology of interfacial Cu6Sn5 IMC in SAC/Cu system during laser soldering. Journal of Materials Science & Technology[J], 2020, 50(0): 115-127 DOI:10.1016/j.jmst.2019.12.036

1. Introduction

Solder joints are utilized in microelectronic and solar photovoltaic (PV) panel manufacturing industries to provide electrical, thermal and mechanical continuities in electronic devices and assemblies. Overall integral performance of devices and circuits is dependent on the strength and reliability of solder joints. The growing popularity of Pb-free solder alloys is attributed to the benefit they provide to the environment and human-health. A large number of lead- free solders have been developed to replace Sn-Pb solders and among them Sn-xAg-yCu (SAC) solders are the most widely studied one. As compared to the traditional Sn-Pb solder alloys, Sn-xAg-yCu solder contains a significantly higher proportion of Sn (usually larger than 90 wt%) [1]. When SAC/Cu system is used for forming solder joints, the larger proportion of Sn in the SAC solder (compared to Sn-Pb system) causes a more pronounced growth of Cu6Sn5 intermetallic compound (IMC) at the interface. As interfacial IMCs are brittle [2], and their occurrence is more prominent in context of SAC solders; any research work studying about interfacial IMC contributes to the understanding of the reliability of solder joints. IMC occupy an increasing proportion of the SAC solder volume in miniaturized Pb-free SAC joints, and thus impact the reliability of the electronic devices [3].

The mechanical behavior of a solder joint is affected by the microstructure and morphology of the interfacial intermetallic compounds (IMCs). The shear strength of a solder joint is dependent on the microstructure not only of the bulk solder matrix but also of the interfacial IMCs. In the shear test performed in the solder joints at a shear height of 10 μm, Yang et al [4] have found out that the shear force for solders with prismatic IMCs is 17.22 N, whereas the shear force for those with scalloped IMCs is 16.03 N. Thus, the solder joints with prismatic IMCs have higher or better shear strength than those with scalloped IMCs.

Laser soldering is a joining technology of solder/substrate characterized by the application of heat only on the desired spot precisely. The moving heat source during laser processing is controlled by a computer and this highly advanced selective soldering can enhance the production rate in joining technology [5]. Especially in context of joint formation at printed circuit boards (PCBs) which consists of many thermally sensitive components, laser soldering has proved to be advantageous over conventional reflow soldering. The attributes such as precisely focused laser beam, localized and non-contact heating, rapid rise and fall in temperature, large cooling rate, short duration etc. make laser soldering a better choice over reflow soldering [[6], [7], [8]]. Nishikawa and Iwata [9] have found out that the solder joints processed by laser has superior impact reliability than those prepared by conventional reflow soldering. Moreover, the growing popularity of laser in flexible electronics [10], additive manufacturing (3D printing) [[11], [12], [13], [14]] and biomaterials fabrication [15] sectors, has attracted a great attention from researchers, and study of laser heat source on any materials system can aid in understanding the mechanism of laser-materials interaction.

The microstructure evolution at the interface of the solder is dependent on the laser-material interaction. The laser processing parameters namely power (P), scan speed (Vscan) etc. and materials composition mainly define the resulting interfacial microstructure of the SAC/Cu system. The transient and spatially non-uniform temperature profiles during laser soldering lead to interfacial microstructural evolution totally different from that of conventional isothermal reflow soldering [16]. The formation of interfacial IMCs of two different morphologies of IMC-prismatic and scalloped, at the interface of Sn/Cu and Sn-3.5Ag-0.5Cu/Cu systems, due to the alteration in the value of laser power and scan speed has been reported in [17]. However, the work uses only two solder compositions and so the solder alloy composition dependence of IMC could not be outlined quantitatively from the study. Provided that solder compositions of more types are introduced for performing laser experiments, a data-driven method can be utilized to infer the exact relationship between the laser parameters, solder composition and the resultant IMC morphology. Lee et al. [13] have utilized data analytics approach for melt-pool geometries in additive manufacturing. The targets for the melt pool geometries (width, depth, area, height etc.) and input features such as powder chemistry, thermal properties, powder size distribution and laser parameters (powder, scan speed, energy density etc.) were obtained from the experiments. In the research of Wang et al. [18] finite element method -generated (FEM-generated) data was utilized for performing the machine learning (ML) task of thermal transport analysis in Aluminum alloys with precipitate morphology.

In this present study, experimental works are performed to observe the morphology of interfacial Cu6Sn5 IMC in Sn-xAg-yCu solders (with 6 different alloy composition features) which are processed at different laser power and scan speed to form solder joints with Cu substrate. In line with the experimental laser power, scan speed and solder composition, finite element analysis is performed to create 51 observations based datasets on laser parameters, solder composition and temperature dependent Jackson parameter (αJ). The fem-obtained datasets are used in a neural network (NN) to predict the IMC morphology based upon the numerical values of αJ. The machine learning based prediction of IMC morphology is validated with experimental results, and critical scan speed required for morphology transition is determined. As this research work integrates together the results from experiments, computational work and data analysis; it can be helpful in the sector of advanced soldering materials design in accordance to the objective of materials genome initiative (MGI) [19].

In SAC solder alloys consisting of Ag, another intermetallic compound, namely Ag3Sn is formed along with Cu6Sn5 compound in the aftermath of laser soldering [17]. Ag3Sn intermetallic nanoparticles are found to be adsorbed on the top of micrometer sized Cu6Sn5 IMC. Although Ag3Sn intermetallic compound influences the growth kinetics of the overall Cu6Sn5 IMC layer, its role on morphology transition of Cu6Sn5 IMC grains is not described adequately in the existing research works. Thus, in this work, the role of Ag3Sn formation with regard to the morphological evolution of Cu6Sn5 IMC grains of SAC solders is assumed negligible. The assessment of Ag3Sn nanoparticles in relation to the morphology of individual Cu6Sn5 IMC grains is a topic beyond the scope of this present work. Similarly, Cu3Sn intermetallic compound is formed between Cu layer and Cu6Sn5 IMC layer only after prolonged aging for several hours at temperature below melting point of solder, and thus Cu3Sn intermetallic formation can be neglected for experimental samples characterized immediately at the aftermath of laser soldering (with total duration of about a few seconds). Thus, only Cu6Sn5 compound is studied as the intermetallic compound at the interface of solder and Cu substrate in the present study. With this consideration, the acronym IMC used in this manuscript, solely represents Cu6Sn5 intermetallic compound and not the other types.

2. Experiments with laser heating

The schematic sketch for the experimental setup for laser soldering is shown in Fig. 1. Pb-free solder sheet with dimensions of 4 mm × 2 mm × 0.2 mm was placed over polycrystalline Cu substrate (5 mm × 5 mm × 0.1 mm). In our previous work [17], only pure Sn and Sn-3.5Ag-0.5Cu solders were considered for study, whereas this works aiming to infer the effect of solder composition of a generic Sn-xAg-yCu solder, additionally considers Sn-0.7Cu, Sn-3.5Ag, Sn-3.0Ag-0.5Cu and Sn-3.5Ag-0.7Cu solders. Thus, the types of solders considered in this study are outlined in Table 1, with their respective Ag and Cu weight fraction based (mi) composition. Since the sum of the weight fraction of n components in a multicomponent system is equal to 1, n-1 components are sufficient to describe the composition of the system. In a ternary Ag-Cu-Sn system, mSn = 1-(mAg+mCu); the weight fractions of Ag and Cu are sufficient to describe the composition of Sn-xAg-yCu solder types.

Fig. 1.

Fig. 1.   Schematic sketch outlining the major variables (P, Vscan and solder composition) in an experimental setup for laser soldering. The point $~\chi$ is located at the bottom corner corresponding to the start point of the laser beam.


Table 1   Composition of Pb-free solder alloys expressed in weight fraction of Ag, Cu and Sn.

Solder typemAgmCumSn
Pure Sn0.00.01.0
Sn-0.7Cu0.00.0070.993
Sn-3.5Ag0.0350.00.965
Sn-3.0Ag-0.5Cu0.030.0050.965
Sn-3.5Ag-0.5Cu0.0350.0050.96
Sn-3.5Ag-0.7Cu0.0350.0070.958

New window| CSV


An SP-50C Fiber laser (wavelength = 1528-1565 nm, pulse frequency = 5000 Hz, focal beam diameter = 25 μm, and pulse width = 2.0 × 10-4 s) was utilized for heating the solder surface. As shown in Fig. 1, the laser beam was aligned perpendicular to the top surface of the Sn-xAg-yCu solder material. Laser output power (P) and scan speed (Vscan) are the two major laser processing parameters, which when varied can influence the materials properties and interfacial microstructure of the SAC/Cu interface. In this study, the values of power were varied as 30 W, 40 W and 50 W. Similarly, depending upon the input power, the magnitudes of scan speed were varied from 10-240 mm/min with a step of 10 mm/min. In order to prevent ablation, scan speed greater than 70 mm/min were used with 50 W power. In the same way, for assurance of joint formation, Vscan ≤ 150 mm/min were used with 30 W power (verification needed). For an example, combination of 50 W and 10 mm/min would lead to the ablation of solder surface whereas 30 W and 240 mm/min could lead to the absence of Cu6Sn5 IMC formation at the interface (no joint formation).

For laser processing of SAC solders with Cu substrate at suitable P and Vscan, the interface melts thereby accelerating the interfacial reaction between the solder and substrate. In such condition, Cu6Sn5 intermetallic compound (IMC) is the predominant one, and for the rest of this work, IMC implies the Cu6Sn5 compound. After the laser soldering of different Sn-xAg-yCu solder types processed with different combinations of P, Vscan, top view morphology of IMC was characterized for observation with scanning electron microscope (SEM). For this purpose, the laser processed samples (joints) were etched in 10% HNO3 solution, and then mounted in epoxy. These samples were polished with SiC sandpapers (different grit sizes), followed by synthetic diamond polishing paste.

Fig. 2 shows the classification of top view morphologies of Cu6Sn5 IMCs in accordance to two stereotypes, namely scalloped and prismatic morphologies for various experimental solder types (Sn, Sn-0.7Cu, Sn-3.5Ag and Sn-3.5Ag-0.7Cu). For the given magnitudes of power (P) and scan speed (Vscan) mentioned in Fig. 2(a), all of the 4 types of solders possess IMCs with scalloped morphology. However, when P and Vscan of values mentioned in Fig. 2(b) are utilized during the laser processing, IMCs of all these solder alloys are found to be of prismatic morphology. Since, larger magnitude of Vscan is responsible for reducing temperature (T) at interface, whereas increase in P is favorable for increasing T (the reason is provided in section: Numerical); these parameters have opposite effects when it comes to IMC morphology. In context of pure Sn solder, Cu6Sn5 IMC has scalloped morphology when laser power is 30 W and scan speed is 40 mm/min. Now, when Vscan = 50 mm/min, the increase of power to 40 W has caused the IMC to bear prismatic morphology. In case of Sn-0.7Cu solder, at P = 30 W; the IMCs possess scalloped and prismatic/faceted morphologies at scan speeds of 70 mm/min and 30 mm/min, respectively. In other words, for 30 W of input laser power, the increase in Vscan from 30 to 70 mm/min caused the transition in IMC morphology from prismatic to scalloped. For Sn-3.5Ag solder alloys, at 30 W and very low scan speed of 20 mm/min, the morphology is prismatic. However, even at a very high magnitude of power (P = 50 W), the IMCs are scalloped at extremely larger scan speed of 220 mm/min. Finally, for Sn-3.5Ag-0.7Cu solder laser processed at 50 W, the IMCs transition from prismatic to scalloped when Vscan is reduced from 130 to 90 mm/min.

Fig. 2.

Fig. 2.   SEM images of experimental solders of different initial composition and laser processed at different combinations of input power (P) and scan speed (Vscan): (a) scalloped and (b) prismatic morphology of Cu6Sn5 IMC compounds as obtained at the interface, respectively.


The intertwined relationship between solder composition and laser processing parameters (P, Vscan) is revealed also in Fig. 3. In case of Fig. 3(a), an experiment performed in Sn-0.7Cu/Cu system at 30 W and 70 mm/min reveals an IMC with scalloped morphology. Now, keeping power constant, if Vscan is reduced to 50 mm/min, the morphology is found to be prismatic. On the other hand, at original scan speed of 70 mm/min and increasing the power to 40 W also makes the morphology appear prismatic. As shown in Fig. 3(b), for laser processing at constant power of 40 W; Sn-3.0Ag-0.5Cu/Cu system produces IMCs with prismatic morphology at Vscan = 100 mm/min, whereas those of Sn/Cu system have a scalloped morphology even at a significantly lower scan speed of 70 mm/min. For experiments performed at constant power of 50 W, IMC corresponding to Vscan = 130 mm/min in case of Sn-3.5Ag-0.7Cu solder alloy possesses scalloped morphology whereas solders such as Sn-3.5Ag and Sn-3.5Ag-0.5Cu have prismatic IMCs even at relatively larger scan speeds of 160 mm/min and 140 mm/min respectively. Thus, Fig. 3(b) suggests that in addition to laser processing parameters, the composition of Sn-xAg-yCu solder has a significant role in deciding the morphology of the resultant IMC during laser soldering of Sn-xAg-yCu/Cu system.

Fig. 3.

Fig. 3.   (a) Effects of input power and scan speed on IMC morphology transition from scalloped to prismatic or vice-versa for a Sn-0.7Cu solder and (b) effects of solder types at given laser input power on IMC morphologies.


3. IMC morphology transition and Jackson parameter

In conventional reflow soldering of a given solder/substrate system, the morphology transition regime for IMCs from scalloped to prismatic or vice-versa can be described by a single magnitude of isothermal reflow temperature. For N experiments of a solder alloy reflowed at m isothermal reflow temperatures, ith temperature is defined as a critical temperature for IMC morphology transition if the IMC bears scalloped morphology below that temperature and prismatic or faceted morphology above that temperature [4]. In case of laser soldering, the temperature in the solder domain is neither uniform nor steady. The spatial and temporal variation in alloy temperature during laser processing, makes it difficult to define a transition temperature. The temperature of the melt pool (near the laser beam spot) is far larger than the distant interface; and even the temperature within the interface is changing with time. So, in this study, instead of reaction temperature, the magnitudes of laser processing parameters (P and Vscan), which influence the temperature of the solder medium, are used as markers for morphology transition. For a given solder composition, and at a given power, the increase in scan speed above a threshold value (Vscan,cr) will cause the morphology transition from prismatic to scalloped. Similarly, for given solder composition and scan speed, an increase in power above the critical value Pcr will cause morphology transition from scalloped to prismatic.

The attainment of a given IMC morphology at a Sn-xAg-yCu/Cu interface can be best defined by Jackson parameter (αJ) [20]. In accordance to the work of Choi et al [21], αJ is an indication of the degree of attachment of atoms to the growing interface, and this parameter predicts whether the interface becomes scalloped (αJ<2) or faceted (αJ>2) In other words, the Cu6Sn5 crystals with αJ<2 are known to have atomically rough interfaces and grow with a non-faceted front, whereas the IMC crystals with αJ>2 possess atomically smooth interfaces with ledges and grow with a faceted front [3]. For the purpose of defining Cu6Sn5 compound, Jackson parameter can defined mathematically by the following expression [4,21]:

${{\alpha }_{J}}=\xi \frac{\Delta {{H}_{ir}}}{R{{T}_{if}}}$

where ΔHir is the enthalpy increment or change during the interfacial reaction of liquid solder and the Cu substrate to form IMC, Tif is the temperature at the interface of liquid solder and Cu during laser soldering, and it is not a constant, R is the universal gas constant, $\xi$ is a fraction of the total number of nearest neighbors in a plane parallel to the considered interface; and this parameter depends on the crystal structure of IMCs and random oriental relationship between adjacent IMC grains. Although there is a common agreement in several research works [21,22] that the value of $\xi$ is always greater than 0 and less than 1, the exact value of this parameter has never been reported. In [4,21,22], mentioning random oriental relationship between evolving IMCs as the major hindrance to calculate the exact value of $\xi$, an average value of 0.5 has been assigned to this parameter.

The enthalpy increment (ΔHir) corresponding to interfacial reaction at solder substrate interface, can be expressed as a polynomial function of Tif [23,24] :

$\Delta {{H}_{\text{ir}}}=A+B{{T}_{\text{if}}}+CT_{\text{if}}^{2}$

Combining Eqs. (1) and (2), Jackson parameter can be rewritten as:

${{\alpha }_{\text{J}}}=\frac{\xi }{R}\left( \frac{A}{{{T}_{\text{if}}}}+B+C{{T}_{\text{if}}} \right)$

Using the experimental values of ΔHir for Cu6Sn5 IMC formation from the works of Choi et al. [21], Flandorfer et al. [25] and Yang et al. [4] have computed the polynomial coefficients A, B, C of Eq. (2) and reported A, B and C to have the values of 2.074 × 104 J/mol, -83.25 J/(mol K) and 0.121 J/(mol K2) respectively. Putting these values in Eq. (3), it can be outlined that for any Tif ≤ 414 K, αJ monotonically increases with temperature (Tif).

4. Numerical determination of interfacial reaction temperature during laser soldering

With the predetermined values of R, A, B, C and $\xi$, Eq. (3), can be utilized to numerically predict the resulting morphology of IMC during laser soldering if Tif is known. Unlike isothermal reflow soldering, the presence of dynamic heat source from laser beam during processing renders the Tif non-uniform and transient. The measurement of such variable is very difficult and one of the possible ways to determine the transient and non-uniform temperature is using finite element analysis (FEA). The numerical results of temperature at a localized region of the interface during the entire timesteps can be averaged to produce the effective temperature denoted as $\overline{{{T}_{\text{if}}}}$.

As shown in Fig. 1, the laser beam is focused on the top surface of the solder alloy, and since the main purpose of numerical analysis is to determine the temperature at the bottom part of the solder (interface between the solder and Cu); a 2D rectangle of length 4 mm and depth (height) 0.2 mm (200 μm) is sufficient to represent the computational domain of the numerical simulation. Thus, the green colored region in Fig. 1, representing the solder alloy is the computational domain of the finite element model. It is assumed that the temperature profiles are symmetrical along the width of solder (that is in the direction perpendicular to laser motion). Triangular mesh elements are used for constructing the finite element mesh.

The partial differential equation (PDE) for phase change based heat transfer during laser soldering in a Sn-xAg-yCu solder [17]:

$\rho {{C}_{\text{eff}}}\left( \frac{\partial T}{\partial t}+\vec{u}\cdot \nabla T \right)={{k}_{\text{th}}}{{\nabla }^{2}}T+{{Q}_{\text{laser}}}$

Where in T is temperature and t is time. In the equation, ρ is the density of the solder alloy. The effective heat capacity (Ceff) of the Sn-xAg-yCu alloy is a function of enthalpy (H) and is described as follows [17]:

${{C}_{\text{eff}}}=\frac{\partial H/\partial t}{\partial T/\partial t}$

The enthalpy in Eq. (5) is defined as $H(T)=\underset{0}{\overset{T}{\mathop \int }}\,\left( \rho {{C}_{\text{p}}}+\rho {{L}_{\text{f}}}\frac{\partial f}{\partial \tau } \right)\text{d}\tau$, where Cp is the specific heat capacity of solder alloy at constant pressure, Lf is the latent heat of fusion of the alloy and f(T) is the temperature dependent liquid-fraction. Enthalpy values for Sn, Sn-0.7Cu, Sn-3.5Ag, Sn-3.0Ag-0.5Cu, Sn-3.5Ag-0.5Cu and Sn-3.5Ag-0.7Cu alloys have been computed as function of T using OpenCalphad software [26,27], and the results are shown in Fig. 4 below. The computed values are in agreement with the experimental results of Morando et al [28]. As depicted in the H-T plot, the vertical lines of the H curves which represent the phase change regime correspond to different T values, implying that the solder alloys have different melting temperatures (Tm). Pure Sn has the highest melting temperature of 505.15 K, whereas Sn-Ag-Cu solders have the lowest melting point of 491 K. Sn-0.7Cu and Sn-3.5Ag alloys are shown in the figure to have vertical lines intermediate between pure Sn and Sn-Ag-Cu solders, and thus have intermediate melting temperatures. This difference in melting temperature is very significant for the temperature evolution during laser heating. It is important to note that the difference in enthalpies measured by the length of the vertical lines of the H-T plot in the two phase regimes, is defined as the latent heat of fusion for a given solder alloy. Within the left hand side of Eq. (4), the second term inside the brackets represents the convection heat transfer (which is more dominant in the region around the melt pool). The fluid flow regime described by velocity (($\vec{u}$) and pressure (p) variables, in the laser melted solder regime is mathematically expressed by the incompressible Navier-Stokes equations assuming Boussinesq approximation, mentioned as following.

$\nabla \cdot \vec{u}=0$
$\rho \frac{\partial \vec{u}}{\partial t}+\rho \left( \vec{u}\cdot \nabla \right)\vec{u}=\nabla \cdot \left[ -pI+\nu \left( \nabla \vec{u}+{{\left( \nabla \vec{u} \right)}^{T}} \right) \right]+{{F}_{\text{boussinesq}}}$

Fig. 4.

Fig. 4.   Enthalpies of Sn-xAg-yCu solder as a function of temperature.


The viscosity of the fluid is denoted by ν in Eq. (7). During laser processing, at a given time (t), the regimes melted by laser heat are in liquid phase, whereas some portion of the solder quite distant may still be in BCT phase, and thermophysical properties for the two phases are quite different. The enthalpy method [29], described by Eq. (4) uses the enthalpy difference between liquid and Body-Centered Tetragonal (BCT) phases, to distinguish the phases in the numerical model. The temperature dependent materials properties can then be implemented separately for the solid BCT and liquid phases. The first term in the right-hand side (RHS) of Eq. (4) is the thermal convection term, and kth is the thermal conductivity of the solder alloy. The second term in the RHS of the heat equation is the heat source term due to laser irradiation, and thus Qlaser is the laser heat generation [30]. Assuming the intensity profile of the laser beam exhibits the Gaussian function, and the laser material interaction is characterized by Beer’s law, Qlaser for the context of 2D simulation, can be expressed by the following equation [31]:

${{Q}_{\text{laser}}}=\frac{2\Omega P}{\pi r_{0}^{2}}{{e}^{(-\frac{2{{r}^{2}}}{r_{0}^{2}}-\Omega \text{z})}}$

where Ω in Eq. (8) is the absorption coefficient of the solder material (μm-1) [32], P is the input power and r0 is the radius of the laser beam. For a laser initially positioned at x0 and moving with a given scan speed (Vscan), the radial distance (r) from the axis of the beam is defined as r = (x0 + tVscan). Now, it is obvious from Eq. (8) that the laser power (P) increases the magnitude of Qlaser, whereas scan speed appearing in the negative exponent term, produces the opposite effect. Subsequently, in accordance to Eq. (4), increasing P helps to raise the medium temperature whereas increasing Vscan will lower the solder temperature. Hence, a combination of the effects of P and Vscan, affect the interfacial reaction temperature (Tif), which will subsequently guide the numerical value of Jackson parameter.

4.1. Material properties

The materials properties used in the energy equation (Eq. 4), continuity and momentum equations (Eqs. (6) and (7)), namely, density, thermal conductivity and viscosity are treated as temperature dependent functions. As defined in [29], ρ and kth are assumed to be have linear dependence on T. The density is assumed to be a continuous function of temperature in both solid and liquid phases. Thermal conductivity function on the other hand, having continuity in each of solid-state temperatures and liquid state temperatures, is treated to have discontinuity in the two phase regime. In accordance to the experimental results of thermal conductivity measurements by Meydaneri et al. [33], the liquid phase in equilibrium with the BCT phase, in the two phase regime, is assigned to have its thermal conductivity 1.10 times higher than that of solid phase even at the same temperature. Arrhenius equation is utilized to describe the temperature dependence of ν in liquid phase, asymptotic viscosity is utilized in mushy region (two-phase region) and BCT phase is described as a region with a very high viscosity [29]. In enthalpy method, which is a typical Eulerian method, the numerical magnitudes of liquid phase viscosity (≈ 1 mPa S) and solid state BCT phase (> 1 Pa S) renders the effective implementation of Navier-Stokes equation in liquid phase only. The values of material properties for the Sn-xAg-yCu solders provided in Table 2, are obtained from [29,[32], [33], [34], [35], [36], [37], [38], [39]]. The thermal expansivity (β) of solders [17], appearing in body force term (Fboussinesq) of Eq. (7), is utilized as a constant in the model. The absorption coefficient (Ω) is taken to be same for all solder alloys, and is chosen for a solder paste containing flux [32].

Table 2   Material properties used in numerical simulation. Thermal conductivity, density and viscosity are utilized as functions of temperature in the finite element model.

PropertiesUnitsPure SnSn-0.7CuSn-3.5AgSn-3.0Ag-0.5CuSn-3.5Ag-0.5CuSn-3.5Ag-0.7Cu
kth (298.15 K)W/(m K)67.053.057.2657.2657.2657.26
ρ(298.15 K)kg/m3731073007360734074507500
ν(823.0 K)mPa S1.1321.5091.1811.4001.4501.550
β(×10-6) K-120.8820.8820.8820.8820.8820.88
Ωμm-17 × 10-37 × 10-37 × 10-37 × 10-37 × 10-37 × 10-3

New window| CSV


4.2. Initial conditions, boundary conditions and numerical implementation

The initial temperature of the computational domain is taken as 298.15 K. The velocity of the medium is taken as zero initially, and pressure is assumed to be equal to that of atmosphere.

Robin boundary conditions (BCs) are imposed for T variable at the boundaries of the numerical domain, in accordance to the following equation:

${{k}_{\text{th}}}\frac{\partial T}{\partial n}={{h}_{\text{n}}}\left( T-{{T}_{\text{ext}}} \right)$

where hn is the convection heat transfer coefficient for a given boundary material n, and Text is the external temperature of the air film. In the top, left and right sides of the domain exposed to air, the, convection heat transfer coefficient (hair) is taken to be equal to 15 W/(m2 K). In the bottom side of the domain touching the Cu substrate, the following time-dependent interfacial heat transfer coefficient (hint) is employed in Eq. (9).

${{h}_{\text{solder}/\text{substrate}}}=\theta {{t}^{-m}}$

Assuming Cu as a moderately smooth surface, amplitude $\theta$ = 10,000 and the time-exponent m = 0.03 [40].

In context of velocity variable, no flux boundary condition (($\frac{\partial \vec{u}}{\partial n}=0$) is chosen for the top side of the solder exposed to air. No slip BC is applied at the bottom side of the solder. With the kinematic condition of no fluid flow across the rigid boundary at the bottom, no slip BC can be mathematically written in the following simplified form:

$\vec{u}={{\vec{u}}_{\text{substrate}}}$

With Cu substrate at rest (u→substrate = 0), Eq. (11) can be rewritten as:

$\vec{u}=0$

The set of Eqs. (4),(6) and (7) along with the provided material properties, initial conditions and boundary conditions, are solved using finite element method (FEM) in Elmer Multiphysics software [[41], [42], [43], [44]]. Since the thickness of the solder is 200 μm and the length is 4 mm; the simulations have been performed at the length scale of 1.0 × 10-4 m; and corresponding to this length scale, the maximum timestep size for Elmer Solver based simulations was kept as 5.0 × 10-2 s. 51 finite element simulations were performed with variation of solder types, P and Vscan. The total or gross simulation time set for the simulations ranged from 1 s (corresponding to largest Vscan = 240 mm/min), and 12.0 s (for the lowest scan speed of 20 mm/min). Using the corresponding solder alloy’s mAg and mCu and applied heat source’s P and Vscan in Eq. (4), a total number of 51 finite element method-based simulations were performed to determine the temperature profile in the solder models.

4.3. Numerical results for thermal simulation under laser irradiation

Fig. 5 shows the temperature distribution in context of pure Sn solder at t = 1.0 s for several combinations of P and Vscan. As shown by the H-T curves for pure Sn in Fig. 4, the melting point of Sn is around 505.15 K, and thus the red colored regions in Fig. 5 located at the vicinity of laser beam spot are all in the molten state. Even the yellow and yellow-green colored portion (corresponding to temperature beyond 505.15 K) is in liquid state. The blue-green and blue colored portion, which denotes T sufficiently less than 500 K, is clearly illustrates that the material is in solid state Body-Centered Tetragonal (BCT) phase. The figure shows that largest mass of liquid solder attains higher temperature (in accordance to the size of red colored regime) in context of laser parameters of 40 W power and 60 mm/min scan speed (case of highest P magnitude). It is followed by the model with P = 30 W and Vscan = 50 mm/min (case of lowest Vscan). The suppressing effect on T raise, due to lowering of power, while keeping the scan speed constant at 60 mm/min can be realized by observing the sizes of red colored regime in the two images- the second image from top (30 W, 60 mm/min) and the bottom-most image (40 W, 60 mm/min). Finally, when the power is kept the lowest at 30 W and scan speed is kept far larger (170 mm/min), where there are no red and yellow colored regimes in the region around the laser beam. A small part of the liquid around the beam spot reaches just above the melting temperature (yellow-green part), and the initial part of the solder is already at solid state.

Fig. 5.

Fig. 5.   Temperature profile corresponding to t = 1.0 s, at the different positions of a pure Sn solder irradiated by different combinations of input power and scan speed.


The interfacial Cu6Sn5 grows at the bottom part of the solder, and the temperature profile (Tif) at the point $\chi$ of Fig. 1 is the property of significant important in this study. Fig. 6 shows the temporal variation of Tif at the bottom point $\chi$ for pure Sn processed with the previously discussed four combinations of P and Vscan. Although the temperature on the top surface at the vicinity of the laser beam is the highest for the case 40 W, 60 mm/min, the time averaged bottom point Tif is the largest for the one with smaller scan velocity (case 30 W, 50 mm/min represented by black colored curved in the figure). Although the Tif values for case 40 W, 60 mm/min (blue colored curve) are initially larger, the larger scan velocity effectively puts the laser beam radially farther from point $\chi$, thus causing a decrease in heating rate in accordance to Eq. (8). The bottom point interface temperature at the same point, is significantly lower in case of laser processing with 30 W, 60 mm/min (red colored curve), but still the solder material reaches the temperature beyond the melting point. With 30 W, 170 mm/min, the average value of Tif is lesser than 400 K, far below the melting temperature of the solder. Although the region around the vicinity of laser beam spot melts, it is clear from the figure that the bottom part of the solder does not reach its melting point. The time-averaged values of bottom corner temperature ((${{\bar{T}}_{\text{if}}}$) for the solder models are 562.73 K (30 W, 50 mm/min), 510.46 K (30 W, 60 mm/min), 389.64 K (30 W, 170 mm/min) and 556.74 K (40 W, 60 mm/min). Fig. 7 presents the temperature distribution at t = 2.0 s in different solder types for P = 30 W, Vscan = 60 mm/min. As the laser beam is already halfway from the initial position at 2 s, the left end extremities of the solders have already cooled down significantly. Thus, to observe the minute differences between the temperature values at the left extremities, the temperature color bar is scaled within the range of 500-600 K for the domain within the first 1 mm from the left and is shown in Fig. 7(a). The temperature profile for the complete length of the solders is exhibited in Fig. 7(b). From Fig. 7(a), it can be understood that extremities temperature is the largest for Sn-0.7Cu solder and the smallest for pure Sn. The Sn-3.5Ag and Sn-3.5Ag0.7Cu solders have the temperature values at the left end (x = 0) intermediate between Sn-0.7Cu and pure Sn alloys. The interfacial region’s temperature i.e. Tif for point χ for these solders is graphically plotted along with time in Fig. 8. It can be noticed in this figure, that for 0.5 s < t < 1.5 s, the temperature at the left bottom corner corresponding to pure Sn solder is in the range 504-506 K (melting point of pure Sn = 505.15 K); whereas in the same duration the temperature values of other solder types are quite higher and they have lower melting point than that of pure Sn. The highest temperature value corresponds to that of Sn-0.7Cu solder, which reaches 550 K at 1.5 s. The time-averaged values of bottom corner temperature ((${{\bar{T}}_{\text{if}}}$) for the different Sn-xAg-yCu solder types at P = 30 W and Vscan = 60 mm/min are 510.46 K (pure Sn), 533.59 K (Sn-0.7Cu), 519.20 K (Sn-3.5Ag) and 519.04 K (Sn-3.5Ag-0.7Cu).

Fig. 6.

Fig. 6.   Numerically computed values of Tif at point $\chi$ in bottom interface for pure Sn modeled with different combinations of P and Vscan. The T-t curves shown in the figure can be utilized for solder joint design for dynamic heat source conditions.


Fig. 7.

Fig. 7.   Temperature profile corresponding to t = 2.0 s, at the different positions of a Sn-xAg-yCu solders irradiated with input power P = 30 W and scan speed = 60 mm/min: (a) the color bar scaled in the range of 500-600 K for getting meticulous temperature profile within the first 1 mm length from the left end; (b) the complete solder geometries and the color bar is set in the range 298-800 K.


Fig. 8.

Fig. 8.   Simulated values of Tif at point $\chi$ in bottom interface for pure Sn-xAg-yCu solder alloys corresponding to P = 30 W and Vscan = 60 mm/min.


5. Influence of ${\bar{T}}_{{if}}$ on interfacial IMC grain growth behavior

As outlined in Section 4.3, the different magnitudes of ${\bar{T}}_{{if}}$ corresponding to laser soldering models of different solder types at same laser processing parameters (P, Vscan) or same solder types subjected to different P and Vscan, would surely impact the rate and behavior of Cu6Sn5 IMC grain growth at the interface. This situation will also naturally imply to different solder types irradiated with varying conditions of P and Vscan. In order to understand how the value of interfacial temperature affects the grain growth behavior of IMCs, a multiphase field simulation [45,46] is performed at two constant interface temperatures Tif of 453.15 K and 523.15 K as shown in Fig. 9, Fig. 10. For the simplification of the model, the content of Ag in the phases is considered to be zero, so that the system can be assumed as a binary Cu-Sn system. The phase field simulation is done using finite element method in Multiphysics Object Oriented Simulation Environment (MOOSE) Framework [47,48]. A rectangle of size 0.17 μm × 0.4 μm is taken as the size of the interface consisting of Sn-rich BCT/Liquid phase (yellowish region), IMC grains 1 and 2 (yellow-green region) and Cu-rich FCC (blue region) phases in Fig. 9. The phases are mathematically represented by non-conserved order parameters ((${{\eta }_{i}}$). The initial size of the identical IMC grains (i.e. at t = 0) are 40 nm × 40 nm (0.040 μm × 40 μm). The FCC phase is considered to have an initial height of 0.15 μm. Parabolic free energy expressions are utilized to describe the bulk free energy of the phases. The values of diffusivities for the phases at 453.15 and 523.15 K are used as provided in [49]. The diffusivity of liquid phase at 523.15 K is 2.0 × 10-12 m2/s whereas that of BCT phase at 453.15 K is 2.0 × 10-16 m2/s. The large difference in the magnitudes of phase diffusivities, corresponding to solder being whether in liquid phase or BCT phase, influences the growth rate of IMC grains. In the figure, the vertical lengths of a grown IMC grain at t = 1.75 h and t = 10 h (Tif =453.15 K) are 0.067 μm and 0.105 μm respectively. On the other hand, the vertical length of grain at t = 0.51 s (Tif = 523.15 K) is about 0.097 μm, and this is just slightly lower than the solid-state grain’s length attained at 10 h. The corresponding length of a grain at t = 1.01 s (Tif = 523.15 K) is 0.153 μm. In the case of grain growth in liquid solder media, since the solid IMC grains provide relative barrier to each other on lateral expansion, the vertical length of two identical grains grows at a faster rate than the lateral growth. In contrary to this, the grain growth rate in solid state solder is nearly equal in both directions. The differential rate of grain growth behavior in lateral and vertical directions for simulation performed at two different temperatures, is even more vividly revealed by the graphical plot of sum of squares of all the order parameters (($\underset{i}{\mathop \sum }\,{{({{\eta }_{i}})}^{2}}$) (shown in Fig. 10). The actual laser soldering process lasts for only several seconds, and for such small-time durations, the IMC grains have practically negligible growth rate at 453.15 K. However, it is noteworthy that number of nearest neighbors around the IMC grain corresponding to t = 0-10 s is far lower for interface temperature of 453.15 K than of interface at T = 523.15 K. This definitely lowers the value of $\xi$ in case of interface with smaller temperature. At t = 4 s (end of laser soldering time), the two grains corresponding to interface at 523.15 K, will almost touch each other laterally and undergo ripening; while growing unperturbed in the vertical direction. This produces compressive stress in one hand, and induces screw dislocation mechanism. Now, at higher interface temperatures, the screw dislocation mechanisms are pronounced to produce faceted or prismatic IMCs upon cooling.

Fig. 9.

Fig. 9.   Result from multi-phase field simulation for grain growth study of two identical IMC grains performed at constant interface temperatures Tif of 453.15 K and 523.15 K.


Fig. 10.

Fig. 10.   Simulated results for sum of squared order parameters (($\underset{i}{\mathop \sum }\,{{({{\eta }_{i}})}^{2}}$) at constant interface temperatures Tif of (a) 453.15 K and (b) 523.15 K.


With the heat source term defined by P and Vscan, laser soldering produces transient temperatures at the interface. Lower scan speed or higher power are associated with overall larger interfacial temperatures, whereas higher scan speed or lower power result into smaller interfacial temperatures. Thus, the results of the phase field simulations performed at isothermal interfacial temperature values can only be utilized for qualitative understanding of IMC grain growth behavior at such interfaces.

6. Prediction of IMC morphology using machine learning

So far now, it has been outlined that the solder alloy composition (mAg, mCu) and the laser processing parameters (P, Vscan) can influence the values of interface temperature in the solder/substrate. FEM-generated datasets have been prepared to correlate mAg, mCu, P and Vscan with Tif. The change in values of Tif directly affects the change in enthalpy (ΔHir) during IMC formation. The use of different solder types may require the use of different $\xi $ in Eq. (1). It can therefore, be understood that αJ is a continuous function of mAg, mCu, P, and Vscan. Since the absolute magnitude of Jackson parameter is an indicator of the morphology of IMC (scalloped or faceted), the mapping of solder composition and laser processing parameters to αJ through machine learning, can subsequently help in quantitative association of alloy composition plus corresponding laser processing parameters to the resulting morphology type.

6.1. Determination of $\xi $ for different solder types

In Eq. (1), αJ is a function of Tif, provided that ξ is determined for a given alloy. The structural parameter can be assumed to be a constant for Cu6Sn5 IMC formation in a given solder alloy. In this study, each solder was assigned constant values ranging from 0.5 to 0.75, with an interval of 0.01. Then the FEM calculated time-averaged ${{\bar{T}}_{\text{if}}}$ for point $~\chi $ at different laser processing parameters were utilized with each magnitude of $~\chi $ for a fixed solder composition, to compute αJ. The optimum $~\chi $ for the given calculations, is the value which yields Jackson parameter perfectly mapped to the experimental morphology. This optimization procedure is repeated to all the solder types. Thus, the value of $~\chi $ are 0.665, 0.734, 0.70, 0.675, 0.670 and 0.524 for pure Sn, Sn-0.7Cu, Sn-3.5Ag, Sn-3.0Ag-0.5Cu, Sn-3.5Ag-0.5Cu, Sn-3.5Ag-0.7Cu solder alloys interfaced with polycrystalline Cu substrate. With the use of fem-generated data and experimental data (scanning electron microscopy images of IMC morphology), the optimum values of $~\chi $ is determined.

6.2. Construction of neural network (NN)

The FEM-generated datasets with 51 observations, are employed to build a neural network (NN). With solder alloy composition variables (mAg, mCu) and laser processing variables (P, Vscan) as input features, a NN shown in Fig. 11 is constructed to estimate αJ as output feature. The hyperparameters of machine learning such as number of layers, number of neurons in each layer, activation function, number of epochs, type of cost function, etc. are then supplied to the model. The NN consists of an input layer, hidden layer(s) and an output layer. The input and output layers have neurons sizes of 4 and 1, respectively. In order to choose the NN with optimized predictability, the number of hidden layers and neurons in them is varied as shown in Table 3 below. As Jackson parameter can be considered to be a continuous function of the input variables, the data analysis task can be considered to be clearly a regression problem. Thus, the NN network is defined to work as a regression-based feed forward neural network. The major benefit of using regression NN is that it can provide fair accuracy even in context of tiny and small datasets [50,51]. Rectified linear unit (ReLU) activation function has been utilized to activate the neurons of a hidden layer or both hidden layers, whereas linear activation function is used at the output layer.

Fig. 11.

Fig. 11.   A feed forward neural network for regression analysis between input features (solder composition and laser processing parameters) and output feature (Jackson parameter).


Table 3   Mean square error at the end of 1000 epochs during neural network analysis with different number of neurons in first hidden layer (HL 1) or two hidden layers. The cases with only one hidden layer and no second hidden layer, are left with blank spaces for the neurons associated with second hidden layer (HL 2).

CasesNeurons in HL 1Neurons in HL 2MSE (Training)MSE (Validation)
I5-2.3 × 10-21.8 × 10-2
II6-1.5 × 10-22.7 × 10-2
III7-1.9 × 10-22.6 × 10-2
IV8-9.0 × 10-32.8 × 10-2
V9-9.8 × 10-31.6 × 10-2
VI10-1.24 × 10-21.96 × 10-2
VII969.7 × 10-31.7 × 10-2
VIII952.0 × 10-43.81 × 10-3
IX852.0 × 10-22.3 × 10-2
X847.0 × 10-31.7 × 10-2

New window| CSV


For an input z, ReLU activation function is defined as f(z) = max{0, z}. In this study, the weight fractions of Ag and Cu considered in the data for solder composition, are well within the range of 0-1. However, laser power is in the range of 30-50. The scan speed ranges from 20 mm/min to 240 mm/min in the datasets. In such contexts, the use of datasets in unprocessed or raw form, may lead Vscan data (having the largest range) to have dominance over the results of neural network analysis. In order to avoid such hegemony of a given input parameter over the result [52,53], and avoid erroneous prediction from the network; a process called normalization can be utilized. Considering that Xi is an element of a column array consisting an input or output parameter of the NN, then normalization of the data is expressed by the following formula:

$\bar{X}=\frac{{{X}_{\text{i}}}-{{X}_{\text{min}}}}{{{X}_{\text{max}}}-{{X}_{\text{min}}}}$

where $\bar{X}$ is the normalized value for a parameter’s element of actual value Xi. Xmax and Xmin are respectively the maximum and minimum values of the parameter. With the normalization procedure via minmax scaler, every elements of mAg, mCu, P, Vscan, predicted αJ (NN-estimated) and expected αJ (FEM-generated), will be within the range of 0-1. In order to convert the predicted αJ in its absolute magnitude, Eq. (13), is utilized again to compute Xi from $\bar{X}$.It is essential to assess the performance of an artificial neural network (ANN) outside the original sample (training datasets) to new datasets (test or validation data). The method of examining the prediction performance of an ANN for a validation data beyond the training data is defined as cross-validation. The original data consisting of 51 observations was split into training and validation datasets in the ratio of 4:1. The mean square error (MSE) between the predicted output (yp) of the network and FEM-computed Jackson parameter (expected αJ) of each dataset [54], is defined as following:

$\text{MSE}=\frac{1}{2n}\sum _{n=1}^{N}{{({{y}_{\text{p}}}-\alpha \text{J})}^{2}}$

Now, this MSE (cost function of the neural network) can be utilized to evaluate the performance (cross-validation) of the ANN first for training and then for validation data. Adam optimizer has been used for the optimization of the NN. With these set of hyperparameters, the neural network model was built, compiled and run in TensorFlow [55] with keras front end [56]. The parameters such as weight matrix ({wi}) and bias matrix ({bi}) are learnt during the model run, and these parameters associate input variables with the output variables of the neural network. The MSE values for training and validation data at the end of 1000 epochs for 10 different NN configuration cases are presented in Table 3. From the table, it is obvious that Case VIII having 9 neurons in the first hidden layer and 5 neurons in the second hidden layer has the least MSE magnitudes of MSE for both the test and validation data. The value of MSE for training data for this neural network (Case VIII) is 2.0 × 10-4, whereas the corresponding validation data has the MSE value of 3.81 × 10-3. Hence, this neural network is selected for further operations on the datasets. For this chosen neural network, the decreasing trend of MSE values of the training and validation data for a total of 1000 epochs is graphically presented in Fig. 12. As the error reduces for both training and validation data, the neural network can be considered to have acceptable performance for further prediction work.

Fig. 12.

Fig. 12.   Mean square error (MSE) values from the neural network analysis of training and validation data for a total of 1000 epochs using 9 neurons in the first hidden layer and 5 neurons in the second hidden layer.


6.3. Using the NN to estimate αJ values of experiments

With successful cross-validation of the neural network, the same NN model is further utilized to estimate the αJ for new datasets consisting of input variables of larger data size. A total of 558 datasets consisting of input variables were employed to perform prediction task. The predicted values of Jackson parameters from the neural network are presented in Fig. 13. As the experimental works have more data of scan speed (ranging from 10 to 240 mm/min) than the laser input power (30-50 W), the scan speed is chosen as a quantity in x-axis of the images of the figure. The numerical values of predicted Jackson parameter are validated with the experimental morphologies corresponding to 51 datasets. The predicted αJ are found to conform with the experimental data (SEM images).

Fig. 13.

Fig. 13.   Jackson parameters estimated from machine learning for different solder alloys at (a) P = 30 W, (b) P = 40 W and (c) P = 50 W. The dotted horizontal lines refer to the morphology transition regimes.


From Fig. 13(a)-(c) it can be inferred that at a fixed laser power magnitude and for a given solder composition, αJ decreases with the increase in scan speed. Moreover, as illustrated in the Fig. 13(a), Sn-0.7Cu solder alloy (red colored curve) has the largest αJ at the vicinity of the dotted horizontal line indicating that it has the highest likelihood to bear prismatic morphology at P = 30 W. However in contexts of P = 40 W (Fig. 13(b)) and P = 50 W (Fig. 13(c)), Sn-3.5Ag solder alloy (green colored curve) has the largest likelihood of bearing prismatic morphology. It can be noticed from Fig. 13(a)-(c), that Jackson parameter for every solder increases with the increment in laser power for a given magnitude of scan speed. For example, in context of Sn-3.5Ag sample processed with Vscan = 20 mm/min, αJ = 2.19 at P = 30 W (Fig. 13(a)), whereas this Jackson parameter has larger magnitudes at higher power values: αJ = 2.668 at P = 40 W (Fig. 13(b)) and αJ = 3.16 at P = 50 W (Fig. 13(c)). The increase in αJ with P, for a given Vscan has a significant implication on the morphological feature of the IMC during laser soldering. IMC transforms into faceted or prismatic grains at αJ > 2, and as scalloped grains at αJ < 2, and this distinction is highlighted as regions separated by a horizontal dotted line corresponding to αJ = 2 in Fig. 13. Thus, at the points of the curves lying above the dotted line IMCs are observed to be bear prismatic morphology whereas they have scalloped morphology at the regions below the dotted line. It is interesting to note that the increment in laser power values also causes an increase in the scan speed values in the x-axes corresponding to the intersection of the dotted lines (αJ = 2) for each Sn-xAg-yCu solder alloy curve. This would help us to understand that at P = 30 W, the interfacial IMCs corresponding to Sn, Sn-3.0Ag-0.5Cu, Sn-3.5Ag-0.5Cu, and Sn-3.5Ag-0.7Cu alloys in Fig. 13(a) are still scalloped for all scan speeds greater than 60 mm/min, whereas at P = 50 W, the IMCs of these 4 solder types have prismatic morphology for Vscan as high as 120 mm/min. The scan speed for a fixed power and solder composition, which acts as a transition boundary between scalloped and prismatic IMC morphologies, has been defined earlier as critical scan speed (Vscan,cr; and thus the increase in solder power will raise the magnitude of the Vscan,cr. The numerical determination of Jackson parameter with the help of neural network analysis, has enabled in numerical quantification of the critical scan speed for a laser processing with given laser parameters and solder/substrate power. In summary, increase in P is characterized with the raise in magnitude of Jackson parameter, whereas increase in Vscan lowers the Jackson parameter.

7. Estimation of critical Vscan,cr for morphology transition

With the success in estimation of numerical value of Jackson parameter through machine learning, it is possible predict the morphology of a resultant interfacial IMC at a solder/substrate interface for given laser power and scan speed. The prediction of IMC morphology is done by classifying the αJ - Vscan curves of Fig. 13 into two regions. As mentioned in the previous section, the classification is done with the help of horizontal line (Jackson parameter = 2). In another way, the classification can be done by a line AB perpendicular to the scan speed axis of Fig. 14, that intersects a Sn-xAg-yCu curve at Jackson parameter = 2. The scan speed corresponding to this line AB that intersects the curves at αJ = 2 is called the critical scan speed. Any points on the curve located in a region at right side of the line has scalloped morphology, whereas that to the left has a prismatic morphology. Also, any point on the curve at a region upper to the horizontal line (αJ = 2) has prismatic morphology, whereas that below the line has scalloped morphology. With these observations, it can be inferred that a point on the curve lying within yellow region of Fig. 14 is characterized with prismatic Cu6Sn5 IMC and a point on the curve within light pink region is featured with scalloped IMC.

Fig. 14.

Fig. 14.   Schematic sketch showing the morphology mapping of IMC in accordance to Jackson parameter. The IMC corresponding to the curve lying in the yellow region has prismatic morphology, whereas Cu6Sn5 within the curve for light pink region has scalloped morphology.


Since scan speed is a laser processing parameter, it is intuitive to utilize critical scan speed (Vscan,cr) as a design parameter for interfacial IMC morphology. The scan speed (Vscan) values, depicted in Fig. 13, have positive values and increase along the positive x-axis. Thus, any scalloped interfacial IMC in the solders lying to the right of line AB. of Fig. 14 has Vscan larger than Vscan,cr whereas any prismatic IMC located to the left of the line has Vscan smaller than Vscan,cr. If Vscan,cr can be predicted for a given solder material processed at a given power, it can be used as a design tool to predict the IMC morphology in advance. In order to estimate the value of Vscan,cr for a given solder material at a given P, the prediction NN network is run by using datasets localized at the vicinity of αJ = 2. For a given solder type at a given power, two input datasets of the previous prediction NN-one yielding αJ just greater than 2 and having least difference with 2.0; and another producing αJ just smaller than 2 and having least difference with 2.0 are selected. These two datasets representing a fixed solder composition and fixed power possess same value of mAg, mCu and P, and they differ by Vscan. Let Vmax,local and Vmin,local be lower and higher limits of scan speed of these chosen two datasets. A total of 100 datasets are constructed and fed into the predictor NN. In each of these 100 datasets mAg, mCu and P do not vary; but Vscan lying between Vmax,local and Vmin,local are changed with a stepping of 0.1 mm/min. The scan speed of the datasets, which predicts the output having the least difference with 2.0 i.e. which corresponds to the smallest value of $\left| {{\alpha }_{\text{J},\text{predict}}}-2 \right|$, is then considered as the optimized value of critical scan speed. The similar process is repeated for different P magnitudes and different solder types.The predicted values of V in different solder types at different laser input power P, are summarized in Fig.15 and Table 4. For all solder types, Vscan,cr is raised with the increase in input laser power, and this increases the probability of occurrence of prismatic IMC morphology at larger laser input power. In case of pure Sn, the critical scan speed at 30 W power is 21.0 mm/min, which means the interfacial IMCs for this condition can retain prismatic morphology only at scan speed lower than 21.0 mm/min. For Sn-3.5Ag/Cu system processed with laser at 50 W, the critical scan speed is 184.3 mm/min, thereby indicating that its interfacial IMCs bear prismatic morphology for all scan speeds lower than 184.3 mm/min; and transform to scallop morphology beyond this critical scan speed. The critical scan speed can be utilized as a reference design parameter to set the power and scan speed parameters for a given alloy, and obtain a desired IMC morphology. From Fig. 15 and Table 4, it can be inferred that Sn-0.7Cu solder has the largest magnitude of critical scan speed (Vscan,cr = 54.4 mm/s) at 30 W input power whereas Sn-3.5Ag alloy has the largest values of critical scan speeds at 40 W and 50 W input power (114.3 mm/min @40 W and 184.3 mm/min @50 W). It implies that these solders retain prismatic IMC morphology even at larger values of scan speed for corresponding power; and thus can be selected as the most preferable solder types in context of applications where joints are exposed to shear loads. The other solders having lower magnitude of critical scan speed are less preferable for applications involving larger shear stress. Since the optimized values of critical scan speed presented in the chart can be utilized for promoting the reliability based design of solder joints in context of shear loads; Fig. 15 can be referred to as a design chart for solders with laser as source of heat.

Fig. 15.

Fig. 15.   Material design chart established from this study that shows the optimized values of Vscan,cr as markers of the Cu6Sn5 IMC morphology transition. For a given laser input power and solder alloy type mentioned in the chart, the IMC has scalloped morphology at Vscan > Vscan,cr, whereas at Vscan < Vscan,cr, it has prismatic morphology.


Table 4   Optimized values of critical scan speed for a solder composition at a given input power (mm/min).

Solder typeP = 30 WP = 40 WP = 50 W
Pure Sn21.067.4141.4
Sn-0.7Cu54.496.7150.0
Sn-3.5Ag44.7114.3184.3
Sn-3.0Ag-0.5Cu38.5102.5172.2
Sn-3.5Ag-0.5Cu33.391.1168.9
Sn-3.5Ag-0.7Cu10.164.0127.9

New window| CSV


Similar, to critical scan speed; another design parameter called critical laser power (Pcr) can be also correlated to Jackson parameter = 2.0. The future works on the determination of critical laser input power, relies on the availability of larger experimental data of input power. Moreover, in future, the effect of other processing parameters of the laser, e.g. wavelength, frequency, beam spot diameter, laser types etc. can be varied additionally in the input datasets to assess their influence on the IMC morphology.

8. Conclusions

This study integrates experiments with finite element analysis and machine learning to design the interfacial Cu6Sn5 IMC at the Sn-xAg-yCu/Cu system. The major conclusions from this research work are presented as following:

(1) Interfacial Cu6Sn5 morphology was investigated for numerous laser processing experiments at Sn-xAg-yCu/Cu interface. The magnitudes of laser processing parameters-power (P) and scan speed (Vscan) and the composition of the solders-weight percentages of Ag (x = mAg) and Cu (y = mCu) are found to influence the morphology of the IMC. The morphology of experimental IMCs affected combinatorically by these variables are classified into two types-prismatic and scalloped.

(2) The Jackson parameter (αJ) is chosen to marker to predict the occurrence of two different morphology. The fact that αJ is the mathematical function of interfacial temperature (Tif) of the solder/substrate interface, it is necessary to associate experimental variables with the interfacial temperature. The solder composition influences the enthalpy and other material properties of the heated system, whereas P and Vscan influence the source term of the heat transfer equation.

(3) A 2D finite element method is developed to simulate the spatial and temporal evolution of temperature for 6 solder types at different power and scan speeds. The temperature field at a point in the bottom interface is time-averaged to obtain average temperature of the solder/substrate interface.

(4) The fem-generated datasets is employed to construct a neural network (NN) model and this regression based NN is used to predict Jackson parameter of an IMC profile from the 4 input variables - namely mAg, mCu, P and Vscan. The condition that prismatic IMCs have αJ > 2 and scalloped ones have αJ < 2 is used for the validation of the prediction results of NN with experimental IMC morphologies. At 30 W, Sn-0.7Cu solder alloy are predicted to have the larger αJ for a given scan speed at the vicinity of morphology transition regime (1.9 < αJ < 2.1). Similarly, at 40 W and 50 W, the Sn-3.5Ag solder alloys is characterized with the largest values of αJ for all values of scan speeds within the regime 1.8 < αJ < 2.2.

(5) The NN predicted αJ values are mapped to the input variables to determine the critical scan speed for a given solder composition exposed to a given laser power. When the laser power is increased, the corresponding critical scan speed for a solder is higher thereby implying the likelihood of prismatic IMC formation. At 30 W input power, Sn-0.7Cu solder has the largest value of critical scan speed (Vscan,cr = 54.4 mm/s). In context of 40 W and 50 W input power, Sn-3.5Ag alloy has critical scan speeds of 114.3 mm/min and 184.3 mm/min respectively, thereby making it the solder type of having the largest Vscan,cr in these two power values. In accordance to this prediction, Sn-0.7Cu/Cu and Sn-3.5Ag/Cu solder joints processed by laser heating in the power range of 30-50 W, are considered the most preferable for applications requiring larger shear strength.

(6) Future work must be creating more experimental data of input laser power, so that the critical power for morphology transition, can also be designed for a broader range of scan speed and solder materials.

Acknowledgements

This work was supported financially by the China Postdoctoral Science Foundation (No. 2017M611215), the “Research Fund for International Young Scientists” of the National Natural Science Foundation of China (No. 51750110504) and the National Natural Science Foundation of China (No. 51871040).

Reference

M.Y. Xiong, L. Zhang, J. Mater. Sci. 54 2019 1741-1768.

DOI      URL     [Cited within: 1]

A. Kunwar, H. Ma, H. Ma, B. Guo, Z. Meng, N. Zhao, M. Huang, J. Mater. Sci. -Mater. El. 27 2016 7699-7706.

DOI      URL     [Cited within: 1]

J.W. Xian, S.A. Belyakov, M. Ollivier, K. Nogita, H. Yasuda, C.M. Gourlay, Acta Mater. 126 2017 540-551.

DOI      URL     [Cited within: 2]

M. Yang, M. Li, L. Wang, Y. Fu, J. Kim, L. Weng, J. Electron. Mater. 40 2011 176-188.

DOI      URL     [Cited within: 5]

N.T. Jaya, S.R.A. Idris, M. Ishak,in: M. Awang (Ed.), The Advances in Joining Technology, Springer, 2019, pp. 97-107.

[Cited within: 1]

H. Nishikawa, N. Iwata, Mater. Trans. 56 2015 1025-1029.

DOI      URL     [Cited within: 1]

T.J. Nabila, S.R.A. Idris, M. Ishak, IOP Conf. Ser. Mater.Sci. Eng. 238 2017, 012011.

[Cited within: 1]

T.J. Nabila, S.R.A. Idris, M. Ishak, IOP Conf. Ser.: Mater.Sci. Eng. 469 2019, 012117.

[Cited within: 1]

H. Nishikawa, N. Iwata, J. Mater. Process. Technol. 215 2015 6-11.

DOI      URL     [Cited within: 1]

J. Bian, L. Zhou, X. Wan, C. Zhu, B. Yang, Y.A. Huang, Adv. Electron. Mater. 5 2019, 1800900.

DOI      URL     [Cited within: 1]

H. Lee, C.H.J. Lim, M.J. Low, N. Tham, V.M. Murukeshan, Y.J. Kim, J. Precis. Eng. Manuf.-Green Technol. 4 2017 307-322.

[Cited within: 1]

S. Wen, K. Chen, W. Li, Y. Zhou, Q. Wei, Y. Shi, Mater. Des. 175 2019, 107811.

DOI      URL     [Cited within: 1]

S. Lee, J. Peng, D. Shin, Y.S. Choi, Sci. Technol. Adv. Mater. 20 2019 972-978.

DOI      URL     PMID      [Cited within: 2]

Modern data analytics was employed to understand and predict physics-based melt-pool formation by fabricating Ni alloy single tracks using powder bed fusion. An extensive database of melt-pool geometries was created, including processing parameters and material characteristics as input features. Correlation analysis provided insight for relationships between process parameters and melt-pools, and enabled the development of meaningful machine learning models via the use of highly correlated features. We successfully demonstrated that data analytics facilitates understanding of the inherent physics and reliable prediction of melt-pool geometries. This approach can serve as a basis for the melt-pool control and process optimization.

E. Fereiduni, A. Ghasemi, M. Elbestawi, Mater. Des. 184 2019, 108185.

DOI      URL     [Cited within: 1]

I. Khadka, Z. Wang, H. Zheng, S. Castagne, J. Micro Nano-Manuf. 7 2019, 024507.

DOI      URL     [Cited within: 1]

T. Hurtony, B. Balogh, P. Gordon, Micro Nanosyst. 2 2010 178-184.

DOI      URL     [Cited within: 1]

A. Kunwar, S. Shang, P. Råback, Y. Wang, J. Givernaud, J. Chen, H. Ma, X. Song, N. Zhao, Microelectron. Reliab. 80 2018 55-67.

DOI      URL     [Cited within: 6]

J. Wang, A.Y. Nobakht, J.D. Blanks, D. Shin, S. Lee, A. Shyam, H. Rezayat, S. Shin, Adv. Theory. Simul. 2 2019, 1800196.

DOI      URL     [Cited within: 1]

J.J. de Pablo, N.E. Jackson, M.A. Webb, L.Q. Chen, J.E. Moore, D. Morgan, R. Jacobs, T. Pollock, D.G. Schlom, E.S. Toberer, J. Analytis, I. Dabo, D.M. DeLongchamp, G.A. Fiete, G.M. Grason, G. Hautier, Y. Mo, K. Rajan, E.J. Reed, E. Rodriguez, V. Stevanovic, J. Suntivich, K. Thornton, J.C. Zhao, npj Comput. Mater. 5 2019 41.

DOI      URL     [Cited within: 1]

K.A. Jackson, Prog. Solid State Chem. 4 1967 53-56.

DOI      URL     [Cited within: 1]

W.K. Choi, S.Y. Jang, J.H. Kim, K.W. Paik, H.M. Lee, J. Mater. Res. 17 2002 597-599.

DOI      URL     [Cited within: 5]

M. Yang, M. Li, C. Wang, Intermetallics 25 (2012) 86-94.

DOI      URL     [Cited within: 2]

The growth behaviors of the intermetallic compounds (IMCs) formed at the eutectic Sn3.5Ag/polycrystalline Cu and pure Sn/polycrystalline Cu interfaces are comparatively studied based on an experiment in which the liquid solder is removed before the end of soldering. This removal of the solder allows for the capture and visualization of the interfacial IMCs formed during liquid-state soldering and avoids the influence of Cu6Sn5 precipitated from the solder matrix during cooling. The results show that round, scallop-type Cu6Sn5 grains with a strong texture form at the molten solder/Cu interface and that their growth is controlled more by grain boundary (GB) diffusion at the beginning of the reaction followed by volume diffusion, whereas the growth of Cu3Sn is only volume-diffusion-controlled. In addition, in contrast to the predictions of some studies, Ag does not inhibit interfacial IMC growth. Instead, by changing the interfacial energy between the molten solder and the interfacial IMC, the addition of Ag affects the growth orientation and coarsening behavior of interfacial Cu6Sn5 grains. These changes lead to more Cu6Sn5 GBs at the interface and therefore greater IMC formation and Cu consumption in the Sn3.5Ag/Cu reaction than in the Sn/Cu reaction under the same reflow conditions. (c) 2012 Elsevier Ltd.

R. Agarwal, Z. Singh, V. Venugopal, J. Alloys. Compd. 282 1999 231-235.

DOI      URL     [Cited within: 1]

D. Chattaraj, R.A. Jat, S.C. Parida, R. Agarwal, S. Dash, Thermochim. Acta 614 (2015) 16-20.

DOI      URL     [Cited within: 1]

H. Flandorfer, U. Saeed, C. Luef, A. Sabbar, H. Ipser, Thermochim. Acta 459 (2007) 34-39.

DOI      URL     [Cited within: 1]

B. Sundman, U.R. Kattner, M. Palumbo, S.G. Fries, Integr. Mater. Manuf. Innov. 4 2015 1-15.

DOI      URL     [Cited within: 1]

B. Sundman, U.R. Kattner, C. Sigli, M. Stratmann, R. Le Tellier, M. Palumbo, S.G. Fries, Comput. Mater. Sci. 125 2016 188-196.

DOI      URL     PMID      [Cited within: 1]

Thermodynamic data are needed for all kinds of simulations of materials processes. Thermodynamics determines the set of stable phases and also provides chemical potentials, compositions and driving forces for nucleation of new phases and phase transformations. Software to simulate materials properties needs accurate and consistent thermodynamic data to predict metastable states that occur during phase transformations. Due to long calculation times thermodynamic data are frequently pre-calculated into

C. Morando, O. Fornaro, O. Garbellini, H. Palacio, J. Mater. Sci. -Mater. El. 25 2014 3440-3447.

DOI      URL     [Cited within: 1]

A. Kunwar, J. Givernaud, H. Ma, Z. Meng, S. Shang, Y. Wang, H. Ma, in: K. Bi, S. Liu, S. Zhou (Eds.), 17th International Conference on Electronic Packaging Technology, ICEPT 2016, IEEE, Wuhan, China, 2016, pp. 166-169.

[Cited within: 3]

F. Caiazzo, V. Alfieri, Materials 11 (2018) 1506.

DOI      URL     [Cited within: 1]

M. Benton, M.R. Hossan, P.R. Konari, S. Gamagedara, MicromachinesBasel 10 (2019) 123.

[Cited within: 1]

C. Hamann, H. Kehrer, H.G. Rosen, C. Scherer, in: W. Waidelich (Ed.), Laser in Engineering, Springer, Berlin, Heidelberg, 1994, pp. 446-452.

[Cited within: 3]

F. Meydaneri, M. Gündüz, M. Özdemir, B. Saatçi, , Met. Mater. Int. 18 2012 77-85.

DOI      URL     [Cited within: 2]

M. Zhao, L. Zhang, Z.Q. Liu, M.Y. Xiong, L. Sun, Sci. Technol. Adv. Mater. 20 2019 421-444.

DOI      URL     PMID      [Cited within: 1]

With the development of lead-free solders in electronic packaging, Sn-Cu lead-free solder has attracted wide attention due to its excellent comprehensive performance and low cost. In this article, we present recent developments in Sn-Cu lead-free solder alloys. From the microstructure and interfacial structure, the evolution law of the internal structure of solder alloy/solder joint was analysed, and the model and theory describing the formation/growth mechanism of interfacial IMC were introduced. In addition, the effects of alloying, particle strengthening and process methods on the properties of Sn-Cu lead-free solders, including wettability, melting and mechanical properties, were described. Finally, we outline the issues that need to be resolved in the future research.

T. Siewert, S. Liu, D.R. Smith, M. Juan, C. Madeni, Database for SolderProperties with Emphasis on New Lead-free Solders, Technical Report, 2002.

[Cited within: 1]

T. Gancarz, P. Fima, J. Pstrus, J. Mater. Eng. Perform. 23 2014 1524-1529.

DOI      URL     [Cited within: 1]

T. Gancarz, Z. Moser, W. Gasior, J. Pstrus, H. Henein, Int. J. Thermophys. 32 2011 1210-1233.

DOI      URL     [Cited within: 1]

This work is aimed at comparing several methods for the measurement of physical properties for molten Sn and Sn-Ag alloys, namely, surface tension, density, and viscosity. The method used for viscosity in this work is the modified capillary method. For surface tension and density, the data used for comparison were previously measured using the maximum bubble pressure method and the dilatometer technique, respectively, for four Sn-Ag alloys having (3.8, 32, 55, and 68) at% Ag. The results are compared with those obtained using a new method based on a fluid draining from a crucible under the influence of gravity, designated the Roach-Henein (RH) method. This new method enables the determination of these three physical properties in one set of measurements. Liquid Sn was used as well as two liquid Sn-Ag alloys having (3.8 and 34.6) at% Ag with the RH method. It was determined that the RH method may be used to simultaneously obtain surface tension, viscosity, and density and that the errors associated with these measurements were similar to those obtained using traditional and separate techniques. Comparisons of the measured viscosity and surface tension to those predicted using thermodynamic models will also be presented. Finally a comparison of mixing model predictions with the experimentally measured alloy surface tension and viscosity is also presented.

W. Gasior, Z. Moser, A. Debski, T. Siewert, Int. J. Thermophys. 31 2010 502-512.

DOI      URL     [Cited within: 1]

www.imim.pl/english) is being integrated with the NIST database (available freely from www.boulder.nist.gov/div853/lead_free/solders.html) to provide a single source for data on lead-free (Pb-free) soldering alloys. The NIST database was developed to support circuit board designers, and so it is focused primarily on mechanical properties. In contrast, the SURDAT database is focused primarily on thermophysical properties. Together, they will form a much more complete source for the properties of Pb-free solder alloys.]]>

M. Tan, B. Xiufang, X. Xianying, Z. Yanning, G. Jing, S. Baoan, Phys. B 387 (2007) 1-5.

DOI      URL     [Cited within: 1]

S. Shang, A. Kunwar, J. Yao, H. Ma, Y. Wang, Thin Solid Films 669 (2019) 198-207.

DOI      URL     [Cited within: 1]

M. Malinen, P. Raback, in: I. Kondov, G. Sutmann (Eds.), Multiscale Modelling Methods for Applications in Materials Science: CECAM Tutorial, 16-20 September 2013, Forschungszentrum Julich GmbH, Julich, Germany, 2013, pp. 101-113.

[Cited within: 1]

A. Kunwar, B. Guo, S. Shang, P. Råback, Y. Wang, J. Chen, H. Ma, X. Song, N. Zhao, Intermetallics 93 (2018) 186-196.

DOI      URL     [Cited within: 1]

B. Guo, A. Kunwar, N. Zhao, J. Chen, Y. Wang, H. Ma, Mater. Res. Bull. 99 2018 239-248.

DOI      URL     [Cited within: 1]

S. Shang, A. Kunwar, J. Yao, Y. Wang, H. Ma, Y. Wang, Met. Mater. Int. 25 2019 499-507.

DOI      URL     [Cited within: 1]

L. Hou, N. Moelans, J. Derakhshandeh, I. De Wolf, E. Beyne, Sci. Rep. 9 2019 14862.

DOI      URL     PMID      [Cited within: 1]

This paper reports on 3D phase field simulations of IMC growth in Co/Sn and Cu/Sn solder systems. In agreement with experimental micrographs, we obtain uniform growth of the CoSn3 phase in Co/Sn solder joints and a non-uniform wavy morphology for the Cu6Sn5 phase in Cu/Sn solder joints. Furthermore, simulations were performed to obtain an insight in the impact of Sn grain size, grain boundary versus bulk diffusion, IMC/Sn interface mobility and Sn grain boundary mobility on IMC morphology and growth kinetics. It is found that grain boundary diffusion in the IMC or Sn phase have a limited impact on the IMC evolution. A wavy IMC morphology is obtained in the simulations when the grain boundary mobility in the Sn phase is relatively large compared to the interface mobility for the IMC/Sn interface, while a uniform IMC morphology is obtained when the Sn grain boundary and IMC/Sn interface mobilities are comparable. For the wavy IMC morphology, a clear effect of the Sn grain size is observed, while for uniform IMC growth, the effect of the Sn grain size is negligible.

J. Hektor, M. Ristinmaa, H. Hallberg, S.A. Hall, S. Iyengar, Acta Mater. 108 2016 98-109.

DOI      URL     [Cited within: 1]

L.K. Aagesen, D. Schwen, K. Ahmed, M.R. Tonks, Comput. Mater. Sci. 140 2017 10-21.

DOI      URL     [Cited within: 1]

S. Shang, Y. Wang, Y. Wang, H. Ma, A. Kunwar, Microelectron. Eng. 208 2019 47-53.

DOI      URL     [Cited within: 1]

M.S. Park, R. Arróyave, Comput. Mater. Sci. 50 2011 1692-1700.

DOI      URL     [Cited within: 1]

D.F. Specht, IEEE Trans. Neural. Netw. 2 1991 568-576.

DOI      URL     PMID      [Cited within: 1]

A memory-based network that provides estimates of continuous variables and converges to the underlying (linear or nonlinear) regression surface is described. The general regression neural network (GRNN) is a one-pass learning algorithm with a highly parallel structure. It is shown that, even with sparse data in a multidimensional measurement space, the algorithm provides smooth transitions from one observed value to another. The algorithmic form can be used for any regression problem in which an assumption of linearity is not justified.

R. Lanouette, J. Thibault, J.L. Valade, Comput. Chem. Eng. 23 1999 1167-1176.

DOI      URL     [Cited within: 1]

S. Fidan, H. Oktay, S. Polat, S. Ozturk, Adv. Mater. Sci. Eng. 2019 (2019), 3831813.

[Cited within: 1]

R. Zhang, J. Li, Q. Li, Y. Qi, Z. Zeng, Y. Qiu, X. Chen, S.K. Kairy, S. Thomas, N. Birbilis, Corros. Sci. 150 2019 268-278.

DOI      URL     [Cited within: 1]

N.H. Christiansen, P.E.T. Voie, O. Winther, J. Høgsberg, J. Appl. Math. 2014 (2014), 759834.

[Cited within: 1]

M. Abadi, A. Agarwal, P. Barham, E. Brevdo, Z. Chen, C. Citro, G.S. Corrado, A. Davis, J. Dean, M. Devin, S. Ghemawat, I. Goodfellow, A. Harp, G. Irving, M. Isard, R. Jozefowicz, Y. Jia, L. Kaiser, M. Kudlur, J. Levenberg, D. Mané, M. Schuster, R. Monga, S. Moore, D. Murray, C. Olah, J. Shlens, B. Steiner, I. Sutskever, K. Talwar, P. Tucker, V. Vanhoucke, V. Vasudevan, F. Viégas, O. Vinyals, P. Warden, M. Wattenberg, M. Wicke, Y. Yu, X. Zheng, TensorFlow: Large-scale Machine Learning on Heterogeneous Systems, 2015, Software available from tensorflow.org.

[Cited within: 1]

Chollet F., et al.: Keras. GitHub (2015). GitHub repository. https://keras.io/getting-started/faq/#how-should-i-cite-keras.

URL     [Cited within: 1]

/