Next Article in Journal
Calculation of and Key Influencing Factors Analysis on Equivalent Resilient Modulus of a Submerged Subgrade
Next Article in Special Issue
Let’s Print an Ecology in 3D (and 4D)
Previous Article in Journal
Research on the Formation Conditions and Preventive Measures of Uranium Precipitates during the Service Process of Medical Isotope Production Reactors
Previous Article in Special Issue
Additive Manufacturing of Fe-Mn-Si-Based Shape Memory Alloys: State of the Art, Challenges and Opportunities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Physics Modeling of Melting-Solidification Characteristics in Laser Powder Bed Fusion Process of 316L Stainless Steel

1
School of Mechanical and Electronic Engineering, Suzhou University, Suzhou 234000, China
2
School of Information Engineering, Suzhou University, Suzhou 234000, China
3
Department of 3D Printing, Korea Institute of Machinery & Materials, Daejeon 34103, Republic of Korea
4
School of Mechanical and Aerospace, Nanyang Technological University, Singapore 639798, Singapore
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(4), 946; https://doi.org/10.3390/ma17040946
Submission received: 26 January 2024 / Revised: 11 February 2024 / Accepted: 16 February 2024 / Published: 18 February 2024
(This article belongs to the Special Issue State of the Art in Materials for Additive Manufacturing)

Abstract

:
In the laser powder bed fusion process, the melting-solidification characteristics of 316L stainless steel have a great effect on the workpiece quality. In this paper, a multi-physics model was constructed using the finite volume method (FVM) to simulate the melting-solidification process of a 316L powder bed via laser powder bed fusion. In this physical model, the phase change process, the influence of temperature gradient on surface tension of molten pool, and the influence of recoil pressure caused by the metal vapor on molten pool surface were considered. Using this model, the effects of laser scanning speed, hatch space, and laser power on temperature distribution, keyhole depth, and workpiece quality were studied. This study can be used to guide the optimization of process parameters, which is beneficial to the improvement of workpiece quality.

1. Introduction

Under the assistance of additive manufacturing (AM) technology, periodic or statistically homogeneous materials can be used to explore a wider range of interesting structures [1]. As a part of the AM technology, laser powder bed fusion (LPBF) is widely used in the manufacturing of metallic material. By using high temperature restricted to local areas, this technology can directly manufacture the metallic powder into any part geometries or complex specimens with some inner degrees of freedom with a layer-by-layer fashion [2,3,4,5]. However, this results in a large temperature gradient, which has great effects on thermal stress. During LPBF, the process parameters have great effects on heat transfer characteristics and molten pool morphology in the melting-solidification process of the powder bed [6,7]. The inappropriate selection of process parameters may also result in undesirable effects, such as thermal cracking, balling, and porosity [8].
The 316L stainless steel (SS316L) has been widely used in navigation, nuclear reactor, aerospace, and medical environments for many years due to its perfect mechanical properties, such as high tensile strength, low coefficient of thermal expansion, perfect resistance to oxidation at high temperature, creep resistance, and fatigue resistance [9,10,11]. In recent years, the SS316L powder has been widely used in AM to create complex shaped structures due to the properties of excellent corrosion resistance, low carbon content, high-temperature resistance, weldability, and good flowability of the powder bed, which makes it well-suited for AM. Moreover, research shows that the nearly full densification part of SS316L can be achieved using the LPBF process [12]. While SS316L offers numerous advantages in AM, there are also some limitations and challenges associated with its use in AM, such as high thermal conductivity, which makes it challenging to achieve optimal melt pool; residual thermal stresses and distortion, which may potentially cause warping or distortion; and inhomogeneous microstructure and porosity, which may affect the mechanical properties. To obtain high-performance final parts of SS316L manufactured by the LPBF process, current research interest in LPBF primarily focuses on optimizing process parameters, analyzing microstructures, characterizing mechanical properties, exploring post-processing techniques, and employing computational modeling and simulation. This collective research effort aims to advance our understanding and enhance the quality of SS316L parts produced via LPBF [13,14,15,16,17,18].
During the LPBF process, the SS316L powder bed is rapidly melted, solidified, and even vaporized. The flow behavior in the molten pool can be affected by these phenomena. Thus, the thermal distribution and molten pool morphology will be changed. To analyze the relationship between the LPBF process and the mechanical properties of final parts, some experimental and theoretical research has been conducted by other scholars. Greco et al. conducted LPBF experiments of SS316L powder with constant laser energy density. Via these experiments, the effects of laser power, layer thickness, and hatch space on workpieces’ roughness, density, and micro-hardness were studied [19]. Röttger et al. manufactured some specimens using SS316L with different types of LPBF devices. Meanwhile, their microstructures were observed by means of scanning electron microscopy. It was found that these specimens possessed similar microstructures, but there was different porosity between the microstructures, which makes the specimens have significant differences in mechanical strength [20]. Suryawanshi et al. compared the part of SS316L produced using LPBF with the conventionally manufactured austenitic stainless steel. By observing their micro and meso-structures, the property variation was analyzed [21]. Şener et al. conducted a series of experiments to study the properties of the samples by changing the process parameters. Via optical profilometry and tensile test, the effects of process parameters on mechanical performance and surface quality were obtained and optimized [22]. The above experiments can efficiently analyze the mechanical properties of the final parts processed by LPBF, but the effect of the melting-solidification mechanism of the metallic powder bed is hard to analyze for the characteristics of molten pool, such as large temperature gradient and flow behavior at a small scale.
Considering the molten pool characteristics, the mechanism model is suitable to analyze the melting-solidification process caused by the rapid temperature change in the LPBF process. The developed numerical model of LPBF by other scholars can be used to guide the improvement of the mechanism model for SS316L. Waqar et al. developed an FE model of LPBF to analyze the influence of process parameters on thermal and stress profile [23]. Trejos et al. also developed an FE model of LPBF, which takes into account most of the manufacturing parameters and thermophysical properties. Using this model, the thermal history of the LPBF process was calculated. By considering nodes with temperatures higher than the melting point as molten status, the evolution of the molten pool dimensions was analyzed [24]. Luo et al. developed a new efficient and accurate FE model for predicting transient material thermal behavior during the LPBF process in part-level. To reduce the computational cost, the adaptive mesh scheme was used, which makes the simulation speed 12× to 18× faster compared to the traditional simulation scheme [25]. As the liquid metal flow behavior was not considered in the above models, it is hard to describe the heat transfer process caused by liquid metal flow. Additionally, due to the wide application of LPBF in industrial production, CAE software (ANSYS version 2019R1) providers also developed solutions to simulate the LPBF process. These solutions can accurately analyze the residual deformations and internal stresses under large temperature gradients [26,27,28], but the inhomogeneity of the powder bed distribution and the liquid metal flow make it hard analyze their effects on the melting-solidification process. To analyze the melting-solidification process of SS316L thoroughly, a comprehensive model coupling heat transfer physics and laminar flow physics is eagerly needed.
In this study, a three-dimensional numerical model of LPBF for SS316L was developed to describe the temperature and velocity field in the melting-solidification process. Via the single and multi-track simulation, the effects of process parameters on temperature distribution, keyhole depth, and workpiece quality were studied. By shedding light on the characteristics of the melting-solidification process, the correlation with the process parameters can be clarified. Based on this study, some phenomena during manufacturing can be explained. Meanwhile, it may contribute to the design of experiments when analyzing the correlation between the process parameters and mechanical properties of 316L stainless steel produced by LPBF.

2. Computational Models

During the LPBF process, complex physical phenomena such as heat transfer, phase change, and fluid flow of liquid metal materials are involved, as shown in Figure 1 [29]. By describing the physical phenomena through the corresponding numerical model, the LPBF process can be quantitatively analyzed, which is of great significance to guide the optimization of workpiece quality.
To analyze the critical physical properties in LPBF, the following assumptions should be made to simplify the coupling physical phenomena.
  • The fluid in the molten pool is assumed to be Newtonian and laminar flow;
  • The mushy zone is assumed to be isotropic during the solid–liquid phase change process;
  • The distribution of powder particles is assumed to be fixed in space;
  • The effect of the surrounding argon’s flow on the molten pool is neglected;
  • The influence of ambient gas on laser energy absorption is neglected.

2.1. Governing Equations

The mass conservation equation in LPBF is as follows [30]:
ρ t + · ρ u = 0
where u is the velocity of liquid metal; ρ is metal density.
For the non-uniform density distribution in the molten pool caused by thermal expansion, the buoyancy effects should be considered. Thus, the momentum conservation equation can be described as follows [31]:
ρ u t + ρ u · u = · μ u p + ρ g + F B F m
where F B is buoyancy described using the Boussinesq approximation F B = ρ g β T T 0 ; β is the thermal expansion coefficient; F m is used to characterize the difference of fluidity caused by the liquid–solid transition, which is calculated as F m = D u ; D is the Darcy drag force term caused by the solidification and can be described as D = C F s 2 / 1 F s 3 + B ; Fs is the local solid fraction in the computational domain; C is a constant that depends on the mushy zone microstructure; B is a small constant that is used to avoid D being infinity when Fs = 1.
The energy conservation equation is given as follows:
t ρ E + · ρ u E = q + · k T
where k is the thermal conductivity; q is the heat source; E is the specific internal energy which can be described as:
E = c T + 1 F s L
where c is the specific heat; L is the latent heat of melting.
To track the free surface of molten pool, the volume of fluid (VOF) method is used, which can be described as:
γ t + · γ u = 0
where γ is the volume fraction of liquid metal material in the computation element.

2.2. Initial and Boundary Conditions

In LPBF, the thermodynamic behavior at the boundary includes a laser heat source, heat dissipation by evaporation, heat convection, and radiation with the surrounding environment. So the thermal boundary condition and initial condition can be defined as follows [32]:
k T n = q x , y , z , t ε r σ s T 4 T 0 4 h c T T 0 q e v
T x , y , z , 0 = T 0
where T0 is the ambient temperature; n is the vector along the normal direction of the top surface of the powder bed; q(x, y, z, t) represents the laser energy distribution; εr is the surface radiation coefficient; σs is the Stefan–Boltzmann constant (5.67 × 10−8 W/(m2·K4)); hc is the natural convection heat transfer coefficient; qev is the heat taken away as the molten pool evaporation.
Based on the characteristics of laser emission, the laser heat source is assumed to be Gaussian distribution [33,34]. The heat flux of the laser decays exponentially from the center outward, and it is given as:
q r = q 0 e 2 r 2 / θ 2
where q is the heat flux; r is the radial distance from the laser center; θ is the laser radius where q = q 0 e 2 ; q0 is the laser beam intensity at r = 0. Thus, q0 can be written as:
q 0 = 2 P ξ π θ 2
where ξ is the absorptivity of the powder bed; P is the laser power. Thus, the heat flux can be rewritten as:
q r = 2 P ξ π θ 2 e 2 r 2 / θ 2
For high laser energy density, the molten pool temperature will exceed the evaporation temperature. The evaporative heat loss can be written as [35]:
q e v = 0.82 Δ H * 2 π M R T P 0 e Δ H * R 1 T v 1 T
where R is the gas constant; P0 is atmospheric pressure; ΔH* is the effective evaporation enthalpy; M is the molar mass; Tv is the evaporation temperature.
As the evaporation of the molten pool under high laser energy density, the recoil pressure is considered as follows [31]:
P r e T = 0.54 P 0 e L e M R 1 T v 1 T
where Le is the latent heat of evaporation; Tv is the evaporation temperature; P0 is the atmospheric pressure.
At the molten pool surface, the pressure balance can be described as [36]:
P 2 μ v n n = P r e σ 1 R 1 + 1 R 2
where n is the normal direction of the molten pool surface; P is the pressure at the molten pool surface; μ is the dynamic viscosity; v n is the normal velocity; R1 and R2 are the principal radii of molten pool surface curvature. The temperature-dependent surface tension coefficient can be described as σ = σ 0 σ / T T T r e f .
For the temperature dependence of the surface tension coefficient, the Marangoni effect appears. The corresponding fluid velocity boundary condition at the molten pool surface can be described as:
μ v t n = σ T T r
where r is the tangential direction of the molten pool surface; v t is the tangential velocity.

2.3. Simulation Cases and Computation

The simulation was carried out using the Flow3D V11.1 commercial software to simulate the melting-solidification process and thermal behavior. The simulation domain size is 1000 μm × 400 μm × 250 μm, and the powder layer thickness is 45 μm. The hexahedral mesh elements were used. The mesh size was set to be 5 μm, which is sufficient to describe the non-uniform powder distribution. The corresponding material properties of SS316L are shown in Table 1, and the temperature-dependent parameters are shown in Figure 2.
Table 1. Thermal and physical properties of SS316L.
Table 1. Thermal and physical properties of SS316L.
Material PropertyValue
Surface tension coefficient1.76 N/m [37]
Change rate of surface tension−4.002 × 10−4 N/m/K [37]
Densitysee Figure 2a
Thermal conductivitysee Figure 2b
Specific heat capacitysee Figure 2c
Solidus temperature1658 K [38]
Liquidus temperature1723 K [38]
Evaporation temperature3086 K [38]
Latent heat of fusion2.8 × 105 J/kg [10]
Latent heat of evaporation7.45 × 106 J/kg [38]
Molar mass5.58 × 10−2 kg/mol [38]
Dynamic viscosity5.6 × 10−3 Pa·s [39]
Surface radiation coefficient0.4 [39]
Absorption coefficient0.55 (solid), 0.3 (liquid) [10]

3. Results and Discussion

3.1. Molten Pool Morphology Analysis

When the molten pool temperature exceeded the evaporation temperature, the recoil pressure induced by the escaping vapor of the liquid metal was larger than its surface tension [41]. Thus, the keyhole formed as shown in Figure 3.
At the center of molten pool, the surface temperature was higher than that at the bottom. Thus, the buoyancy appeared due to thermal expansion, resulting in natural convection flowing from the bottom up. The temperature of the molten pool center was also higher than that at the edge. Thus, the surface tension at the molten pool center was smaller than that at the edge for the negative change rate of surface tension. Under the effects of the surface tension, the liquid metal at the center was drawn to the edge [42]. The corresponding flow pattern is shown in Figure 3. This outward flow can promote the bonding of the adjacent scan track after solidification. The asymmetrical circulation in the molten pool was caused by the inhomogeneity of powder distribution, which also created the rough contour line of the scan track after the molten pool solidification.

3.2. Temperature Distribution Analysis

Linear energy density (LED) is commonly used to analyze the temperature distribution in LPBF [43]. Figure 4 shows the temperature distribution on the substrate along the scanning direction under the LED of 240 J/m, 400 J/m, 600 J/m, and 800 J/m respectively, which was obtained at the same laser power of 120 W. It can be found that when the molten pool temperature exceeded the evaporation temperature, the variation of LED had little effect on the temperature near the laser focus. This is because the input energy was taken away by evaporation [41], but the keyhole width increased significantly with the increase of LED for the evaporation enhancement. When the LED decreased to 240 J/m, the recoil pressure on the molten pool surface (Knudsen layer) was wakened, and the keyhole depth decreased significantly, which made it so the laser energy could not be sufficiently transferred to the substrate. Thus, the substrate temperature was lower than the liquidus temperature, which means the previously processed layer and the currently processed one could not be effectively fused in the product manufacturing process, resulting in poor densification [8]. If the LED was further decreased, the densification would drop sharply due to the incomplete melting of the powders [44].
To analyze the substrate temperature at laser focus, a temperature sampling probe was set on the substrate, which moved with the laser focus synchronously. To well observe the temperature on the substrate, the keyhole formation on the substrate should be avoided. Thus, the laser power = 120 W and scanning speed = 0.45 m/s were used in this simulation. The temperature at the probe is shown in Figure 5. It could be found that when the laser turned on, the temperature rose quickly. At t = 0.15 ms, the probe temperature reached the liquidus temperature, and the molten pool formed. For the energy absorption of powder melting, the probe temperature of the molten pool rose slowly until the molten pool formed sufficiently. Then, the probe temperature fluctuated around 2635 K (dashed line) due to the non-uniform powder distribution.
Moreover, a fixed temperature sampling probe was set up on the substrate to analyze the multi-track scanning process. In this simulation, the laser power = 120 W, scanning speed = 0.45 m/s, and hatch space = 0.12 mm. The corresponding scanning strategy is shown in Figure 6, and the evolution of temperature is shown in Figure 7. It can be seen that the sampling temperature decreased periodically. In each period, the temperature rose rapidly when the laser focus was close to the sampling probe, and the temperature decreased relatively slowly when the laser focus was far away from the sampling probe. The peak temperature in each period was extracted and fitted as the dashed line in Figure 7, which can be described using the following formula:
T t = 2777 e 0.2931 t + 1212 e 0.0024 t
where t(ms) is laser scanning time. Based on this equation, it can be predicted that the temperature can be reduced to 320 K when the scanning time is 556 ms.

3.3. Keyhole Depth Analysis

The keyhole has great effects on the pore formation, which may degrade the performance of the final parts [45]. To simulate the keyhole formation process, different laser scanning speeds and scanning powers were analyzed. Figure 8 shows the molten pool surface distribution for four different scanning speeds (0.15, 0.2, 0.3, and 0.4 m/s) for the same laser power of 120 W, respectively. Figure 9 shows the molten pool surface distribution for five different laser powers (90, 105, 120, 135, and 150 W) for the same scanning speed of 0.2 m/s, respectively. Then, the keyhole depth was obtained, as shown in Figure 10 and Figure 11, respectively. It could be found that with the decrease of scanning speed and the increase of laser power, the gradient of the keyhole depth also increased. This indicates that the keyhole depth can be increased more effectively by increasing the energy input after the keyhole has formed. Moreover, the remelting of the previous track can increase the densification and wetting properties [44].
To analyze the multiple-track interaction, the multiple-track manufacturing process was simulated with laser power = 120 W, scanning speed = 0.2 m/s, and hatch space = 0.12 mm. The molten pool surface distribution of the second and third tracks is shown in Figure 12. It shows that the two molten pools had the same morphology, which means that the residual heat from the previous track had little effect on the keyhole depth of the current one.

3.4. Hatch Space Analysis

When the laser was 90 W, the scanning speed was 0.2 m/s, the layer thickness was 45 μm, and the stable scan track width was 0.12 mm, as shown in Figure 13a. When the hatch space was set as 0.14 mm, which is larger than the stable scan track width, the internal void defects formed between the adjacent scan tracks, as shown in Figure 13b. To avoid the internal void defects, more overlap of the scan track was needed. Thus, the hatch space was reduced to 0.12 mm, and the stable scan track width was 0.25 mm between adjacent scan tracks, as shown in Figure 13c. By lowering the hatch space, the parts with better density and ultimate tensile strength are manufactured [46].

3.5. Analysis of Powder Bed Distribution Effects

When the powders are distributed sparsely in one location, the molten pool morphology will be affected. If the sparse region is located at the boundary of the scan track, as shown in Figure 14a, then the boundary roughness of the solidified region will be worse, as shown in Figure 14b, which is not beneficial to enlarge the scan track width.
If the sparse region is located at the center of the scan track, as shown in Figure 14c, this will result in the internal void defects, as shown in Figure 14d (scanning speed = 0.2 m/s, laser power = 120 W). For such void defects, it can be avoided by increasing the laser power to delay the solidification of the molten pool, which can promote the release of air bubbles in the molten pool, as shown in Figure 14e (scanning speed = 0.2 m/s, laser power = 150 W).

4. Conclusions

In this study, the multi-physics simulation model was developed to analyze the melting-solidification process of SS316L in the laser powder bed fusion. Using this model, the temperature distribution and molten pool morphology were analyzed by changing the process parameters. Moreover, the effects of powder distribution on workpiece quality were analyzed. Some valuable conclusions were drawn, as follows:
  • The convection flow in the molten pool can effectively widen the molten pool width and promote strong bonding between adjacent scan tracks. Therefore, the hatch space can be enlarged by increasing the laser power or decreasing the scanning speed to enhance the convection flow behavior.
  • When the LED decreases to 240 J/m, the keyhole depth becomes too small to fuse the previously processed layer with the currently processed one, potentially leading to degradation in part densification. To ensure better densification of the final parts, it is suggested that the LED be set to over 400 J/m when the layer thickness is 45 μm.
  • The keyhole depth can be enlarged more effectively by further increasing the energy input after the keyhole is formed, such as by increasing the laser power or decreasing the scanning speed.
  • If the hatch space is too large or the powder bed is sparsely distributed, internal void defects may form, significantly affecting workpiece quality. To prevent these defects, it is suggested that the hatch space be narrower than the single-track width.
Furthermore, the comprehensive liquid–gas phase change model was not considered for the obstacle of integration with the current simulation model. This may result in the deviation of the temperature field between our numerical simulation and experiments when reaching evaporation temperature. In the follow-up research, a sophisticated simulation model including the liquid–gas phase change should be developed to comprehensively analyze the dynamics of the temperature field after the appearance of evaporation.

Author Contributions

Conceptualization, X.S. and H.Z.; methodology, J.-P.C. and H.Z.; software, X.S.; validation, Z.P.; formal analysis, L.H.; investigation, L.H.; resources, L.H.; data curation, M.G.; writing—original draft preparation, X.S.; writing—review and editing, J.-P.C. and H.Z.; visualization, Z.P.; supervision, X.S.; project administration, X.S.; funding acquisition, J.-P.C. and H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Research Project of Education Department of Anhui Province, grant number 2022AH051378, 2023AH052230 and 2023AH052221; the Key Research and Technology Development Projects of Anhui Province, grant number 202004a06020045; the Doctor of Suzhou University Scientific Research Foundation Project, grant number 2021BSK001 and 2021BSK023; Scientific Research platform of Suzhou University, grant number 2021XJPT25ZC, 2021XJPT24ZC and 2021XJPT51; and the Research and Development Fund Project of Suzhou University, grant number 2021fzjj29. This research was also supported by the Basic Research Program funded by the Korea Institute of Machinery and Materials (KIMM), grant number NK248I, and the Technology Innovation Program funded by the Ministry of Trade, Industry, and Energy (MOTIE, Korea), grant number 20024344.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Milton, G.; Briane, M.; Harutyunyan, D. On the possible effective elasticity tensors of 2-dimensional and 3-dimensional printed materials. Math. Mech. Complex Syst. 2017, 5, 41–94. [Google Scholar] [CrossRef]
  2. DebRoy, T.; Wei, H.L.; Zuback, J.S.; Mukherjee, T.; Elmer, J.W.; Milewski, J.O.; Beese, A.M.; Wilson-Heid, A.; De, A.; Zhang, W. Additive manufacturing of metallic components—Process, structure and properties. Prog. Mater. Sci. 2018, 92, 112–224. [Google Scholar]
  3. Gao, B.W.; Zhao, H.J.; Peng, L.Q.; Sun, Z.X. A Review of Research Progress in Selective Laser Melting (SLM). Micromachines 2023, 14, 57. [Google Scholar] [CrossRef]
  4. Sefene, E.M. State-of-the-art of selective laser melting process: A comprehensive review. J. Manuf. Syst. 2022, 63, 250–274. [Google Scholar] [CrossRef]
  5. Ciallella, A.; Giorgio, I.; Barchiesi, E.; Alaimo, G.; Cattenone, A.; Smaniotto, B.; Vintache, A.; D’Annibale, F.; dell’Isola, F.; Hild, F.; et al. A 3D pantographic metamaterial behaving as a mechanical shield: Experimental and numerical evidence. Mater. Des. 2024, 237, 112554. [Google Scholar] [CrossRef]
  6. Wang, Z.; Xiao, Z.; Tse, Y.; Huang, C.; Zhang, W. Optimization of processing parameters and establishment of a relationship between microstructure and mechanical properties of SLM titanium alloy. Opt. Laser Technol. 2019, 112, 159–167. [Google Scholar] [CrossRef]
  7. Lo, Y.-L.; Liu, B.-Y.; Tran, H.-C. Optimized hatch space selection in double-scanning track selective laser melting process. Int. J. Adv. Manuf. Technol. 2019, 105, 2989–3006. [Google Scholar] [CrossRef]
  8. Chowdhury, S.; Yadaiah, N.; Prakash, C.; Ramakrishna, S.; Dixit, S.; Gulta, L.R.; Buddhi, D. Laser powder bed fusion: A state-of-the-art review of the technology, materials, properties & defects, and numerical modelling. J. Mater. Res. Technol.-JmrT 2022, 20, 2109–2172. [Google Scholar]
  9. Chen, J.; Wei, H.; Bao, K.; Zhang, X.; Cao, Y.; Peng, Y.; Kong, J.; Wang, K. Dynamic mechanical properties of 316L stainless steel fabricated by an additive manufacturing process. J. Mater. Res. Technol. 2021, 11, 170–179. [Google Scholar] [CrossRef]
  10. Ansari, P.; Rehman, A.U.; Pitir, F.; Veziroglu, S.; Mishra, Y.K.; Aktas, O.C.; Salamci, M.U. Selective Laser Melting of 316L Austenitic Stainless Steel: Detailed Process Understanding Using Multiphysics Simulation and Experimentation. Metals 2021, 11, 1076. [Google Scholar] [CrossRef]
  11. Wang, D.; Wu, S.; Yang, Y.; Dou, W.; Deng, S.; Wang, Z.; Li, S. The Effect of a Scanning Strategy on the Residual Stress of 316L Steel Parts Fabricated by Selective Laser Melting (SLM). Materials 2018, 11, 1821. [Google Scholar] [CrossRef]
  12. Miranda, G.; Faria, S.; Bartolomeu, F.; Pinto, E.; Madeira, S.; Mateus, A.; Carreira, P.; Alves, N.; Silva, F.S.; Carvalho, O. Predictive models for physical and mechanical properties of 316L stainless steel produced by selective laser melting. Mater. Sci. Eng. A 2016, 657, 43–56. [Google Scholar] [CrossRef]
  13. Ponticelli, G.S.; Venettacci, S.; Giannini, O.; Guarino, S.; Horn, M. Fuzzy process optimization of laser powder bed fusion of 316L stainless steel. Prog. Addit. Manuf. 2022, 8, 437–458. [Google Scholar] [CrossRef]
  14. Theeda, S.; Jagdale, S.H.; Ravichander, B.B.; Kumar, G. Optimization of Process Parameters in Laser Powder Bed Fusion of SS 316L Parts Using Artificial Neural Networks. Metals 2023, 13, 842. [Google Scholar] [CrossRef]
  15. Qiu, C.; Kindi, M.A.; Aladawi, A.S.; Hatmi, I.A. A comprehensive study on microstructure and tensile behaviour of a selectively laser melted stainless steel. Sci. Rep. 2018, 8, 7785. [Google Scholar] [CrossRef]
  16. Iqbal, N.; Jimenez-Melero, E.; Ankalkhope, U.; Lawrence, J. Microstructure and Mechanical Properties of 316L Stainless Steel Fabricated Using Selective Laser Melting. MRS Adv. 2019, 4, 2431–2439. [Google Scholar] [CrossRef]
  17. Kaynak, Y.; Kitay, O. The effect of post-processing operations on surface characteristics of 316L stainless steel produced by selective laser melting. Addit. Manuf. 2019, 26, 84–93. [Google Scholar] [CrossRef]
  18. Soylemez, E.; Koç, E.; Coşkun, M. Thermo-mechanical simulations of selective laser melting for AlSi10Mg alloy to predict the part-scale deformations. Prog. Addit. Manuf. 2019, 4, 465–478. [Google Scholar] [CrossRef]
  19. Greco, S.; Gutzeit, K.; Hotz, H.; Kirsch, B.; Aurich, J.C. Selective laser melting (SLM) of AISI 316L—Impact of laser power, layer thickness, and hatch spacing on roughness, density, and microhardness at constant input energy density. Int. J. Adv. Manuf. Technol. 2020, 108, 1551–1562. [Google Scholar] [CrossRef]
  20. Röttger, A.; Boes, J.; Theisen, W.; Thiele, M.; Esen, C.; Edelmann, A.; Hellmann, R. Microstructure and mechanical properties of 316L austenitic stainless steel processed by different SLM devices. Int. J. Adv. Manuf. Technol. 2020, 108, 769–783. [Google Scholar] [CrossRef]
  21. Suryawanshi, J.; Prashanth, K.G.; Ramamurty, U. Mechanical behavior of selective laser melted 316L stainless steel. Mater. Sci. Eng. A 2017, 696, 113–121. [Google Scholar] [CrossRef]
  22. Şener, B.; Çavuşoğlu, O.; Yuce, C. Effects of process parameters on surface quality and mechanical performance of 316L stainless steel produced by selective laser melting. Optik 2023, 287, 171050. [Google Scholar] [CrossRef]
  23. Waqar, S.; Guo, K.; Sun, J. FEM analysis of thermal and residual stress profile in selective laser melting of 316L stainless steel. J. Manuf. Process. 2021, 66, 81–100. [Google Scholar] [CrossRef]
  24. Trejos-Taborda, J.; Reyes-Osorio, L.; Garza, C.; del Carmen Zambrano-Robledo, P.; Lopez-Botello, O. Finite element modeling of melt pool dynamics in laser powder bed fusion of 316L stainless steel. Int. J. Adv. Manuf. Technol. 2022, 120, 3947–3961. [Google Scholar] [CrossRef]
  25. Luo, Z.; Zhao, Y. Efficient thermal finite element modeling of selective laser melting of Inconel 718. Comput. Mech. 2019, 65, 763–787. [Google Scholar] [CrossRef]
  26. Mayer, T.; Brandle, G.; Schonenberger, A.; Eberlein, R. Simulation and validation of residual deformations in additive manufacturing of metal parts. Heliyon 2020, 6, e03987. [Google Scholar] [CrossRef] [PubMed]
  27. Pagac, M.; Hajnys, J.; Halama, R.; Aldabash, T.; Mesicek, J.; Jancar, L.; Jansa, J. Prediction of Model Distortion by FEM in 3D Printing via the Selective Laser Melting of Stainless Steel AISI 316L. Appl. Sci. 2021, 11, 1656. [Google Scholar] [CrossRef]
  28. Kumar, A.; Shukla, M. Numerical Modeling of Selective Laser Melting: Influence of Process Parameters on the Melt Pool Geometry. J. Mater. Eng. Perform. 2022, 32, 7998–8013. [Google Scholar] [CrossRef]
  29. Ahmed, N.; Barsoum, I.; Haidemenopoulos, G.; Al-Rub, R.K.A. Process parameter selection and optimization of laser powder bed fusion for 316L stainless steel: A review. J. Manuf. Process. 2022, 75, 415–434. [Google Scholar] [CrossRef]
  30. Gan, Z.; Yu, G.; He, X.; Li, S. Numerical simulation of thermal behavior and multicomponent mass transfer in direct laser deposition of Co-base alloy on steel. Int. J. Heat Mass Transf. 2017, 104, 28–38. [Google Scholar] [CrossRef]
  31. Rasch, M.; Roider, C.; Kohl, S.; Strauß, J.; Maurer, N.; Nagulin, K.Y.; Schmidt, M. Shaped laser beam profiles for heat conduction welding of aluminium-copper alloys. Opt. Lasers Eng. 2019, 115, 179–189. [Google Scholar] [CrossRef]
  32. Luo, C.; Qiu, J.; Yan, Y.; Yang, J.; Uher, C.; Tang, X. Finite element analysis of temperature and stress fields during the selective laser melting process of thermoelectric SnTe. J. Mater. Process. Technol. 2018, 261, 74–85. [Google Scholar] [CrossRef]
  33. Hussein, A.; Hao, L.; Yan, C.; Everson, R. Finite element simulation of the temperature and stress fields in single layers built without-support in selective laser melting. Mater. Des. (1980–2015) 2013, 52, 638–647. [Google Scholar] [CrossRef]
  34. Shi, R.; Khairallah, S.A.; Roehling, T.T.; Heo, T.W.; McKeown, J.T.; Matthews, M.J. Microstructural control in metal laser powder bed fusion additive manufacturing using laser beam shaping strategy. Acta Mater. 2020, 184, 284–305. [Google Scholar] [CrossRef]
  35. Guo, Z.; Wang, L.; Wang, C.; Ding, X.; Liu, J. Heat Transfer, Molten Pool Flow Micro-Simulation, and Experimental Research on Molybdenum Alloys Fabricated via Selective Laser Melting. Materials 2020, 14, 75. [Google Scholar] [CrossRef]
  36. Cho, M.H.; Lim, Y.C.; Farson, D.F. Simulation of weld pool dynamics in the stationary pulsed gas metal arc welding process and final weld shape. Weld. J. 2006, 85, 271–283. [Google Scholar]
  37. Cao, L. Workpiece-scale numerical simulations of SLM molten pool dynamic behavior of 316L stainless steel. Comput. Math. Appl. 2021, 96, 209–228. [Google Scholar] [CrossRef]
  38. Ebrahimi, A.; Sattari, M.; Bremer, S.J.L.; Luckabauer, M.; Römer, G.-w.R.B.E.; Richardson, I.M.; Kleijn, C.R.; Hermans, M.J.M. The influence of laser characteristics on internal flow behaviour in laser melting of metallic substrates. Mater. Des. 2022, 214, 110385. [Google Scholar] [CrossRef]
  39. Chang, C.-S.; Wu, K.-T.; Han, C.-F.; Tsai, T.-W.; Liu, S.-H.; Lin, J.-F. Establishment of the Model Widely Valid for the Melting and Vaporization Zones in Selective Laser Melting Printings Via Experimental Verifications. Int. J. Precis. Eng. Manuf.-Green Technol. 2021, 9, 143–162. [Google Scholar] [CrossRef]
  40. Kim, C.S. Thermophysical Properties of Stainless Steels; Argonne National Lab.: Lemont, IL, USA, 1975. [Google Scholar] [CrossRef]
  41. Fürstenau, J.-P.; Wessels, H.; Weißenfels, C.; Wriggers, P. Generating virtual process maps of SLM using powder-scale SPH simulations. Comput. Part. Mech. 2019, 7, 655–677. [Google Scholar] [CrossRef]
  42. Le, T.-N.; Lo, Y.-L. Effects of sulfur concentration and Marangoni convection on melt-pool formation in transition mode of selective laser melting process. Mater. Des. 2019, 179, 107866. [Google Scholar] [CrossRef]
  43. Mao, S.; Zhang, D.Z.; Ren, Z.; Fu, G.; Ma, X. Effects of process parameters on interfacial characterization and mechanical properties of 316L/CuCrZr functionally graded material by selective laser melting. J. Alloys Compd. 2022, 899, 163256. [Google Scholar] [CrossRef]
  44. Wei, K.; Wang, Z.; Zeng, X. Influence of element vaporization on formability, composition, microstructure, and mechanical performance of the selective laser melted Mg–Zn–Zr components. Mater. Lett. 2015, 156, 187–190. [Google Scholar] [CrossRef]
  45. Shrestha, S.; Starr, T.; Chou, K. A Study of Keyhole Porosity in Selective Laser Melting: Single-Track Scanning with Micro-CT Analysis. J. Manuf. Sci. Eng. 2019, 141, 071004. [Google Scholar] [CrossRef]
  46. Huang, M.; Zhang, Z.; Chen, P. Effect of selective laser melting process parameters on microstructure and mechanical properties of 316L stainless steel helical micro-diameter spring. Int. J. Adv. Manuf. Technol. 2019, 104, 2117–2131. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the LPBF process. The gray represents the unprocessed steel; the blue represents the molten pool; and the baby blue represents the molten pool after solidification.
Figure 1. Schematic diagram of the LPBF process. The gray represents the unprocessed steel; the blue represents the molten pool; and the baby blue represents the molten pool after solidification.
Materials 17 00946 g001
Figure 2. Temperature-dependent parameters of SS316L: (a) density [38]; (b) thermal conductivity [40]; (c) specific heat capacity [40].
Figure 2. Temperature-dependent parameters of SS316L: (a) density [38]; (b) thermal conductivity [40]; (c) specific heat capacity [40].
Materials 17 00946 g002
Figure 3. Flow behavior in the molten pool with different cross sections (laser power = 120 W, scanning speed = 0.2 m/s). The black arrows show the fluid flow; the arrow length shows the velocity magnitude; and the color scale shows the temperature field in Kelvin.
Figure 3. Flow behavior in the molten pool with different cross sections (laser power = 120 W, scanning speed = 0.2 m/s). The black arrows show the fluid flow; the arrow length shows the velocity magnitude; and the color scale shows the temperature field in Kelvin.
Materials 17 00946 g003
Figure 4. Temperature distribution under different LEDs.
Figure 4. Temperature distribution under different LEDs.
Materials 17 00946 g004
Figure 5. Substrate temperature at laser focus.
Figure 5. Substrate temperature at laser focus.
Materials 17 00946 g005
Figure 6. Scanning strategy.
Figure 6. Scanning strategy.
Materials 17 00946 g006
Figure 7. Evolution of substrate temperature at the fixed sampling probe.
Figure 7. Evolution of substrate temperature at the fixed sampling probe.
Materials 17 00946 g007
Figure 8. Molten pool surface distribution under different scanning speeds.
Figure 8. Molten pool surface distribution under different scanning speeds.
Materials 17 00946 g008
Figure 9. Molten pool surface distribution under different scanning powers.
Figure 9. Molten pool surface distribution under different scanning powers.
Materials 17 00946 g009
Figure 10. Keyhole depth under different scanning speeds.
Figure 10. Keyhole depth under different scanning speeds.
Materials 17 00946 g010
Figure 11. Keyhole depth under different scanning powers.
Figure 11. Keyhole depth under different scanning powers.
Materials 17 00946 g011
Figure 12. Molten pool surface distribution during multiple-track fabrication.
Figure 12. Molten pool surface distribution during multiple-track fabrication.
Materials 17 00946 g012
Figure 13. Cross section of molten pool after solidification under different hatch space: (a) single-track scanning; (b) hatch space = 0.14 mm; (c) hatch space = 0.12 mm.
Figure 13. Cross section of molten pool after solidification under different hatch space: (a) single-track scanning; (b) hatch space = 0.14 mm; (c) hatch space = 0.12 mm.
Materials 17 00946 g013
Figure 14. The effects of the powder bed distribution: (a) sparse region at the boundary of scan track; (b) defect region caused by the sparse region in (a); (c) sparse region at the center of the scan track; (d) defect region caused by the sparse region in (c); (e) normal morphology of the scan track after increasing the laser power.
Figure 14. The effects of the powder bed distribution: (a) sparse region at the boundary of scan track; (b) defect region caused by the sparse region in (a); (c) sparse region at the center of the scan track; (d) defect region caused by the sparse region in (c); (e) normal morphology of the scan track after increasing the laser power.
Materials 17 00946 g014
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shan, X.; Pan, Z.; Gao, M.; Han, L.; Choi, J.-P.; Zhang, H. Multi-Physics Modeling of Melting-Solidification Characteristics in Laser Powder Bed Fusion Process of 316L Stainless Steel. Materials 2024, 17, 946. https://doi.org/10.3390/ma17040946

AMA Style

Shan X, Pan Z, Gao M, Han L, Choi J-P, Zhang H. Multi-Physics Modeling of Melting-Solidification Characteristics in Laser Powder Bed Fusion Process of 316L Stainless Steel. Materials. 2024; 17(4):946. https://doi.org/10.3390/ma17040946

Chicago/Turabian Style

Shan, Xiuyang, Zhenggao Pan, Mengdi Gao, Lu Han, Joon-Phil Choi, and Haining Zhang. 2024. "Multi-Physics Modeling of Melting-Solidification Characteristics in Laser Powder Bed Fusion Process of 316L Stainless Steel" Materials 17, no. 4: 946. https://doi.org/10.3390/ma17040946

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

Article Metrics

Back to TopTop