1 Introduction

Nickel-based superalloys are used extensively in gas turbine engine applications owing to their unique combination of excellent properties including high tensile strength, low oxidation and good creep resistance at elevated temperature for long exposure times. CMSX-4 is a second-generation, rhenium-bearing nickel-based single crystal superalloy with ultrahigh strength, particularly with regard to long-term creep behavior.[1,2] Recognition that the presence of grain boundaries accelerates creep led to the development of single crystal blades for the hot stage of gas turbine engines, thereby contributing to elevated operating temperatures and higher efficiency. Such components are subject to wear during service which means that there has been sustained work to develop repair methods that restore the integrity of the component and maintain the single crystal microstructure.[3,4,5]

Before the advent of additive manufacturing (AM), repair attempts on single crystal superalloys were performed with a variety of welding processes. Severe defects such as cracks[6,7] and stray grains[8,9] frequently occur during this repair. The formation of new grains with crystallographic orientations different from that of the base material during solidification, commonly referred to as “stray grains”, is of significant interest due to its deleterious effects during the repair of nickel-based single crystal superalloys.[3,10] As a result, the single crystal structure of the material is lost and the mechanical properties are compromised in comparison to the original component. Such stray grains correspond to the onset of equiaxed solidification that replaces epitaxial growth under certain conditions.

To mitigate stray grain formation, previous work has focused on understanding the dendritic solidification behavior and the processing conditions that affect the stray grain formation during solidification within the fusion zone (FZ).[11,12,13,14] Researchers have proposed several possible mechanisms for stray grain formation during welding of single crystal alloys.[12,13,14,15] The localized constitutional undercooling ahead of the solidification front can lead to heterogeneous nucleation and growth of equiaxed grains. In addition, dendrite fragmentation caused by vigorous fluid flow in the melt pool can also lead to the formation of new grains near the melt pool boundary. In both mechanisms, stray grain formation relies on nucleation sites, the difference being that dendrite fragmentation implies that pieces of dendrites act as the nucleation sites whereas the other mechanism uses any other type of nucleation site found in the material, e.g., oxide particles. As is well-known, the converse approach for many castings is to promote equiaxed solidification via the introduction of nucleants such as TiB\(_2\) in aluminum alloys.

In the constitutional undercooling mechanism, Hunt[11] developed a model to describe the columnar to equiaxed transition (CET) under steady state conditions. Gaumann and Kurz modified Hunt’s model and explained the extent to which stray grains can nucleate and grow during the solidification of single crystals.[12,14] In later research, Vitek improved on Gaumann’s model and included a more detailed analysis of the effect of weld conditions, such as power and scanning velocity. Vitek also included the effect of substrate orientation on the stray grain formation through both experimental and modeling techniques.[3,10] It was found that, in general, high weld speeds and low powers minimize the amount of stray grains and maximize the epitaxial single crystal growth during laser welding process.[3,10] Vitek did not, however, consider dendrite fragmentation and his work indicated that heterogeneous nucleation is the primary mechanism leading to stray grain formation in laser welded CMSX-4 single crystal alloy. In the present work, Vitek’s numerical methods are employed and extended to the higher speeds and lower powers characteristic of the rapid nature of metals AM.

The manufacturing of metal components via AM has witnessed a drastic increase in popularity over the last decade.[16] Though the possibility of fabricating CMSX-4 by electron beam melting (EBM) has been frequently investigated,[17,18,19,20,21] a very limited number of investigations have been conducted on the manufacture and repair of CMSX-4 single crystal components using laser powder bed fusion (LPBF), a popular subset of AM, especially as it pertains to the mechanism of mitigating stray grain formation.[22] This lack of investigation is primarily caused by the processing challenges related to these alloy systems.[2,19,22,23,24] Strict control of the processing parameters (e.g., the heat source power, scanning velocity, spot size, preheating temperature, and scan strategy) is essential to be able to build fully dense parts and maintain a single crystal microstructure.[25] While crack-free repair of single crystal alloys is currently possible using EBM,[19,24] repairs that do not generate stray grains are not easily achievable.[23,26]

In this work, laser melting with conditions representative of LPBF was employed to investigate stray grain mitigation in single crystal CMSX-4. LPBF uses a scanning laser beam to melt and fuse a thin layer of metallic powder to a substrate.[27] The use of a laser beam in the layer-wise deposition generates a melt pool that experiences steep temperature gradients, rapid heating/cooling cycles, and violent fluid flow. This usually leads to highly dynamic physical phenomena that can cause defects in the parts.[28,29,30] The dynamics of laser-induced keyhole (i.e., a topological depression caused by vaporization-induced recoil pressure) and thermofluidic flows are strongly coupled and associated with solidification defects in AM processes.[31,32,33,34] An unstable keyhole that is prone to sudden changes of geometry can lead to detrimental physical defects which include porosity, balling, spatter formation, and uncommon microstructure phases.[35] However, keyhole evolution and fluid flow are naturally difficult to capture via traditional post-mortem characterization techniques. To take advantage of high-fidelity numerical modeling, computational fluid dynamics (CFD) was applied in this study to clarify the laser–matter interaction beneath the surface.[36] This brings keyhole and melt pool fluid flow quantification into practice as compared to long-established post-mortem characterization of the cross-sections of solidified melt pools.

In order to develop a proper procedure for the laser-based AM repair and manufacture of CMSX-4 components, we start by establishing a suitable process window and develo** a predictive capability for the propensity of stray grain formation during solidification. Single-track deposition on multiple alloys has been shown to provide an accurate analysis of melt pool geometry and microstructure in AM processes with or without a powder layer.[37,38,39] Accordingly, this study used single-track laser scan experiments without powder to examine the solidification behavior of CMSX-4. This provides a preliminary experimental guideline for LPBF manufacture of CMSX-4 single crystal. In addition, the solidification modeling was extended from traditional welding to the rapid welding associated with LPBF to identify optimal laser melting conditions for the reduction of stray grains. The highly dynamic fluid flow in the melt pool was modeled to provide further guidance for the processing parameter optimization.

2 Materials and Methods

2.1 Single-Track Experiments

Samples were fabricated from a CMSX-4 directionally solidified single crystal ingot using electrical discharge machining (EDM). The final geometry of the samples were in the form of cuboids of dimensions 20 \(\times \) 20 \(\times \) 6 mm. One of the six \(\langle 001\rangle \) crystallographic directions of the ingot was positioned normal to the cutting surface to ensure that the laser tracks scanned along this preferred growth direction. Single laser-melted tracks were made on a powder-free surface of the sample using an EOS M290 machine. This machine is a LPBF system equipped with an ytterbium fiber laser with maximum power output of 400 W and Gaussian beam diameter of 100 \(\mu \)m at the focal spot. During the experiment, the rectangular samples were fitted into pockets in a customized sample holder for the LPBF machine to keep the surfaces at the same height. More details of this customized sample holder are described elsewhere.[39] Experiments were conducted under an argon purged atmosphere and no preheating was applied. Single-track laser melting experiments were performed under various laser powers (200 to 370 W) and scan speeds (0.4 to 1.4 m/s).

2.2 Characterization

After laser scanning, the samples were sectioned using a diamond saw through the FZ on the plane perpendicular to the laser beam scanning direction. Subsequently, the samples were mounted and auto-ground starting with 220 grit SiC papers and concluding with a colloidal silica suspension polish. Crystallographic characterization was carried out on a TESCAN MIRA 3XMH field emission scanning electron microscope (SEM) at an accelerating voltage of 20 kV. EBSD maps were acquired at a \(0.4\,\mu {\text {m}}\) step size. EBSD data were cleaned up and analyzed using Bruker system. The EBSD cleanup began with a grain dilation routine to bring the grains into contact and then followed with a neighbor orientation cleanup routine to resolve the black pixels associated with unindexed diffraction patterns. The cross-sections were also analyzed under an optical microscope to analyze the melt pool morphology. To enhance the contrast in optical characterization, samples were etched with a variation of Marbles reagent composed of 10 g CuSO\(_4\), 50 mL HCl and 70 mL H\(_2\)O.

2.3 Solidification Modeling

Solidification modeling based on the constitutional undercooling criterion was carried out to evaluate the effect of processing parameters on the propensity and distribution of stray grains. Details of this analytical modeling approach are provided in previous work.[3,10] As stated in Reference 3, the variation of total stray grain area fraction in a melt pool with crystallographic orientation of the base material is minimal, therefore, the effect of base material orientation was not considered in this work. In order to compare our LPBF results with the previous work, the mathematically simple Rosenthal equation used in Vitek’s work[3] was also adopted here as a baseline to calculate the shape of melt pool and the thermal conditions in the FZ as a function of the laser parameters. The Rosenthal solution assumes heat is transmitted via conduction only through a steady state point source on a semi-infinite plate with constant material properties and is typically expressed as follows[40,41]:

$$\begin{aligned} T = T_0 + \frac{\eta P}{2\pi k\sqrt{x^2 + y^2 + z^2}} \exp \left[ \frac{-V(\sqrt{x^2 + y^2 + z^2}-x)}{2\alpha } \right] , \end{aligned}$$
(1)

where T is the temperature, \({T_0}\) is the ambient temperature that is set to 313 K (i.e., EOS machine chamber temperature) in this study, P is laser beam power, V is laser beam scanning velocity, \(\eta \) is laser absorptivity, k is thermal conductivity, and \(\alpha \) is thermal diffusivity of the base alloy. The x, y, and z are directions aligned with the opposite of the laser scanning direction, the transverse direction and the vertical direction, respectively. This orthogonal coordinate followed the system in Figure 1 in Reference 3. The solidification front (i.e., solid–liquid interface) was defined as the isotherm average of the solidus temperature (1603 K) and liquidus temperature (1669 K) for CMSX-4.[42,43,44] Thermophysical properties used in the simulation are listed in Table I.

Table I Thermophysical Properties Used for Solidification Modeling of CMSX-4

The thermal gradient was determined by the external heat flow \(\nabla T\) as given by[45]:

$$\begin{aligned} G = |\nabla T| = \left| \frac{\partial T}{\partial x}{\hat{i}}+\frac{\partial T}{\partial y}{\hat{j}} + \frac{\partial T}{\partial z}{\hat{k}} \right| = \sqrt{\left( \frac{\partial T}{\partial x} \right) ^2 + \left( \frac{\partial T}{\partial y}\right) ^2 +\left( \frac{\partial T}{\partial z}\right) ^2}, \end{aligned}$$
(2)

where \({\hat{i}}\), \({\hat{j}}\), and \({\hat{k}}\) are unit vectors along the x, y, and z directions, respectively. The solidification isotherm velocity, \(v_T\), is geometrically related to the laser beam scanning velocity, V, by the following relation:

$$\begin{aligned} v_T = V\cos {\theta } = V \frac{\frac{\partial T}{\partial x}}{\sqrt{\left( \frac{\partial T}{\partial x} \right) ^2 + \left( \frac{\partial T}{\partial y}\right) ^2+ \left( \frac{\partial T}{\partial z}\right) ^2} }, \end{aligned}$$
(3)

where \(\theta \) is the angle between the scanning direction and normal direction of the solidification front (i.e., the direction of maximum heat flow). In constrained growth, such as the welding conditions in this study, the dendritic solidification front is forced to grow at the velocity of the solid–liquid isotherm, \(v_T\).[46]

The local area fraction of newly nucleated grains ahead of the advancing solidification front \(\Phi \), is determined by the liquid temperature gradient G, solidification front velocity \(v_T\), and the nucleus density \(N_0\). By assuming all grains nucleate at a fixed critical undercooling \(\Delta T_n\), the radius of the equiaxed grains is obtained by integrating the growth velocity from the time that the grains start to nucleate to the time the columnar fronts reach the grains. Substituting time by undercooling \({\text {d}}(\Delta T)/{\text {d}}t = -v_T G\), the following relationship between the thermal gradient G, the local volume fraction of equiaxed grains \(\Phi \), the dendrite tip undercooling \(\Delta T\), the nucleus density \(N_0\), the material parameter n, and the nucleation undercooling \(\Delta T_n\), was derived by Gäumann et al.[12,14] based on a modification of Hunt’s model[11]:

$$\begin{aligned} G = \frac{1}{n+1}\root 3 \of {\frac{-4\pi N_0}{3\ln (1-\Phi )}} \Delta T \left( 1- \frac{\Delta T_n^{n+1}}{\Delta T^{n+1}}\right) . \end{aligned}$$
(4)

To simplify the calculation we approximate the dendrite tip undercooling to be solely that of the constitutional undercooling, \(\Delta T_{\text {c}}\), which can be approximated by the power law form \(\Delta T_{\text {c}} = (av_T)^{1/n}\), where a and n are material-dependent constants. For CMSX-4, these values are \(a = 1.25\times 10^6\) s K3.4 m\(^{-1}\), \(n=3.4\), and \(N_0 = 2\times 10^{15}\) m\(^{-3},\) as reported by Reference 3. \(\Delta T_n\) is 2.5 K and can be neglected for solidification under cooling rates greater than \(10^6\) K/s. The expression for \(\Phi \) is solved by rearranging the above equation:

$$\begin{aligned} \begin{aligned} \Phi&= 1-e^S \\&\text {\ where\ } S =\frac{-4\pi N_0}{3}\left( \frac{1}{(n+1)(G^n/av_T)^{1/n}} \right) ^3 = -2.356\times 10^{19}\left( \frac{v_T}{G^{3.4}}\right) ^{\frac{3}{3.4}}. \end{aligned} \end{aligned}$$
(5)

As proposed by Hunt,[11] a value of \(\Phi \le 0.66\) pct represents fully columnar epitaxial growth condition, and, conversely, a value of \(\Phi \ge 49\) pct indicates that the initial single crystal microstructure is fully replaced by an equiaxed microstructure. To calculate the overall stray grain area fraction, we followed Vitek’s method by dividing the FZ into roughly 19 to 28 discrete parts (depending on the length of the melt pool) of equal length from the point of maximum width to the end of melt pool along the x direction. The values of G and \(v_T\) were determined at the center on the melt pool boundary of each section and these values were used to represent the entire section. The area-weighted average of \(\Phi \) over these discrete sections along the length of melt pool is designated as \({\overline{\Phi }}\), and is given by:

$$\begin{aligned} {\overline{\Phi }} = \frac{\sum _k A_k {\Phi }_k }{\sum _k A_k}, \end{aligned}$$
(6)

where k is the index for each subsection, and \(A_k\) and \(\Phi _k\) are the areas and \(\Phi \) values for each subsection. The summation is taken over all the sections along the melt pool. Vitek’s improved model allows the calculation of stray grain area fraction by considering the melt pool geometry and variations of G and \(v_T\) around the tail end of the pool.

Over the years, many advanced numerical methods have been developed to improve the accuracy of modeling melt pool phenomena. We used high-fidelity CFD with FLOW-3D, which is a commercial finite volume method (FVM) that incorporates multiple physics models.[47,48] CFD numerically simulates fluid motion and heat transfer and the primary physics models used here were the laser and surface force models. In the laser model, multiple reflections and Fresnel absorption are implemented through the ray-tracing technique.[36] First, the laser beam is discretized into multiple rays based on each grid cell that is illuminated by the laser beam. Then for each incident ray, some portion of its energy is absorbed by the metal when the incident vector is aligned with the normal vector of the metal surface at the incident location. The absorptivity is estimated using the Fresnel equation.[36] The rest of the energy is retained by the reflected ray, which is treated as a new incident ray if it hits a material surface. Two major forces act on the surface of the liquid metal and deform the free surface. Recoil pressure created by the evaporation of the metal is the primary force causing vapor depression. The recoil pressure model used in this study is \(P_{\text {r}}=A \exp \{ B(1-T_v/T) \}\), where \(P_{\text {r}}\) is recoil pressure, A and B are coefficients related to material properties, which are 75 and 15, respectively. \(T_v\) is the saturation temperature and T is the temperature of the keyhole wall. The other driving force for surface flow and keyhole formation is surface tension. The surface tension coefficient is estimated as a linear function of temperature to include Marangoni flow, \(\sigma = 1.79 - 9.90 \cdot 10^{-4} (T - 1654\,{\text {K}})\) N m\(^{-1}\).[49] The computational domain is one-half of a bare plate (2300 \(\mu \)m \(\times \) 250 \(\mu \)m \(\times \) 500 \(\mu \)m) with a symmetric boundary condition applied on the xz plane. The mesh size is 8 \(\mu \)m and the time step is 0.15 \(\mu \)s, which provides a balance between computational efficiency and accuracy.

3 Results and Discussion

3.1 Melt Pool Morphology

The five laser powers (P) and six scanning velocities (V), utilized in this work generated 29 melt pools with different \(P{-}V\) combinations. The one with highest P and V values was not analyzed in this study because it is associated with excessive balling based on Figure 1.

The single-track melt pools, as shown in Figure 1, can be categorized into four types based on the geometry[39]: (1) conduction mode (in blue boxes), (2) keyhole mode (red), (3) transition mode (magenta), (4) balling mode (green). In keyhole mode, with the typical combination of high laser power and low scanning velocity, the melt pool usually presents a deep and slender shape, with width/depth (W/D) ratios much greater than 0.5. As the scanning velocity increases, the melt pool becomes shallower, presenting a semicircular conduction mode melt pool where W/D is approximately 0.5. The W/D for transition mode melt pools is between 1 and 0.5. A further increase of the scanning velocity to 1200 and 1400 mm/s can generate a sufficiently large cap height and excessive undercutting, which is characteristic of a balling mode melt pool.

Melt pool depth and width as a function of power and velocity are plotted in Figures 2(a) and (b), respectively. The melt pool width was measured at the substrate surface. Figure 2(a) shows that the depth follows a highly linear relationship with laser power. As the velocity increases, the slope of the depth vs. power curves decreases steadily, although there is some overlap in the higher velocity curves. This unexpected overlap may be due to the effects of fluid flow that often causes dynamic variation in melt pool morphology and the fact that only one image was extracted per laser scan. Such linear behavior is not as obvious for width in Figure 2(b). Figure 2(c) shows the melt depth and width as a function of linear energy density P/V. The linear energy density is a measure of the energy input per unit length of the deposit.[3 exhibit a better single crystalline quality as evidenced by the absence of stray grains. It is obvious that minimal grain nucleation occurs in these shallower melt pools created by a scanning laser with relatively low power and high velocity. Stray grains with larger area fractions are observed more frequently in deep melt pools created under high power and low velocity. The effects of power and velocity on the local solidification conditions will be investigated in the subsequent modeling section.

Fig. 3
figure 3

EBSD inverse pole figure maps (color-coded with respect to the x axis direction) and corresponding inverse pole figures showing crystal orientations in solidified melt pools processed by various laser power and velocity conditions (a) P = 200 W, V = 600 mm/s, (b) P = 200 W, V = 1000 mm/s, (c) P = 200 W, V = 1400 mm/s, (d) P = 250 W, V = 600 mm/s, (e) P = 250 W, V = 1000 mm/s, (f) P = 250 W, V = 1400 mm/s, (g) P = 300 W, V = 600 mm/s, (h) P = 300 W, V = 1000 mm/s, and (i) P = 300 W, V = 1400 mm/s. The black dashed curves delineate melt pool boundaries. The scale bar is equivalent to 50 \(\mu \)m in all subfigures. A 5 deg misorientation is used as the threshold to differentiate stray grains (Color figure online)

3.4 Solidification Modeling

As stated in the introduction, researchers evaluated possible mechanisms of stray grain formation during welding of single crystal.[12,13,14,15,55] The two most popular mechanisms discussed are (1) heterogeneous nucleation aided by constitutional undercooling ahead of the solidification front and (2) dendrite fragmentation caused by fluid flow in the melt pool. The first mechanism has been studied extensively. Taking a binary alloy as an example, the solid cannot accommodate as much solute as the liquid, so it rejects solute into the liquid during solidification. Consequently, solute partitioning ahead of the growing dendrites creates an undercooled liquid of which the actual temperature is lower than the local equilibrium liquidus. The existence of a sufficiently extensive constitutionally undercooled zone promotes nucleation and growth of new grains.[56] Note that the total undercooling is the sum of several contributions including constitutional, kinetic, and curvature undercooling at the solidification front. A common assumption is that the kinetic and curvature undercooling can be neglected with respect to the larger contribution of the solutal undercooling for alloys.[57]

To better understand the underlying mechanisms at different \(P{-}V\) conditions, solidification modeling is carried out. The first purpose is to evaluate the overall extent of stray grains (\({\overline{\Phi }}\)) as a function of processing parameters and to examine the variation of local stray grain fraction (\(\Phi \)) as a function of location in the melt pool. The second purpose is to understand the relationship between the solidification microstructure and the stray grain formation mechanism during the fast solidification of metal AM.

Fig. 4
figure 4

Plot of analytical simulated area-weighted stray grain fraction \({\overline{\Phi }}\) vs. laser scanning velocity for three different powers (a) P = 200 W, (b) P=250 W, and (c) P = 300 W, as practiced in LPBF using varying laser absorptivity values

Figure 4 displays the analytically simulated stray grain fraction \({\overline{\Phi }}\) for various laser scanning velocity and laser powers under three laser absorptivity values. The results reveal that the stray grain area fraction is sensitive to the absorbed energy. By increasing the absorptivity from 0.30 to 0.80, the value of \({\overline{\Phi }}\) is approximately tripled, and this effect is more pronounced in low velocity and high power regions. All else being equal, the large impact of absorbed power is attributed a general decrease in average thermal gradient magnitude and increase in average solidification rate within the melt pool. The average stray grain fraction decreases as the scanning velocity increases and power decreases. This general trend is in line with the simulated welding result in blue region of Figure 5, adapted from Vitek’s work.[3] The larger undercooled zone (i.e., \(G/v_T\) zone) in a conventional weld pool means that the area fraction of stray grains in a weld pool is orders of magnitude larger than that for LPBF conditions corresponding to the pink region. In spite of this, the general tendencies of both data sets are similar, i.e., the fraction of stray grains decreases as the laser power decreases and laser velocity increases. It can also be inferred from Figure 5 that the effect of variation in laser parameters on the stray grain area fraction diminishes as the scanning velocity increases into the LPBF region. Figure 6(a) compares the experimental stray grain area fraction from EBSD analysis in Figure 3 with analytical simulation results in Figure 4. Although the exact values differ in keyhole shaped FZs, the trend was consistent in both simulation and experimental data. The keyhole shaped melt pools-especially the two with power 300 W have a significantly larger amount of stray grains than the analytically simulation prediction. This discrepancy is actually expected as the Rosenthal equation usually fails to properly reflect the heat flow in the keyhole regime due to its assumption that heat transfer is governed purely by conduction.[39,40] It also further validates the findings from Figure 4, namely that the increase in absorbed power during keyhole modes leads to more ideal conditions for stray grain formation. Figure 6(b) compares the experimental \({\overline{\Phi }}\) with numerical CFD simulated \({\overline{\Phi }}\). Although the CFD model slightly over predicts \({\overline{\Phi }}\) in all \(P{-}V\) conditions, its prediction under keyhole condition is more accurate than the analytical model. For conduction mode melt pools, the experimental values align more closely with the analytical simulated values.

Fig. 5
figure 5

Plot of analytical simulated area-weighted stray grain fraction \({\overline{\Phi }}\) vs. laser moving velocity for laser welding conditions at the green region which is adapted from Ref. 3, and laser melting practiced in LPBF using an absorptivity 0.8 in the pink region (Color figure online)

Inspection on the distribution of the simulated temperature gradient G and solidification rate \(v_T\) pairs from analytical modeling is shown on the CMSX-4 microstructure selection map in Figure 7(a). Provided that \(G/v_T\) (i.e., morphology factor) controls the morphology and \(G \times v_T\) (i.e., cooling rate) controls the scale of the solidified microstructure,[3