1 Introduction
316L is austenitic stainless steel with low carbon content and is broadly used in biomedical, chemical, petrochemical, transportation industries, and nuclear power plants [1-3]. Some of the medical and industrial components made of 316L are dental implants, orthopedic screws, bone plates, engine parts, pipes, heat exchangers, pumps, axes, and so on [4-6]. Although the biocompatibility, high strength, good ductility, corrosion resistance, and low-cost characteristics of 316L make it a good candidate for medical components and industrial applications [3, 7, 8], its low tribological properties such as low wear resistance, low surface hardness, and high friction coefficient restrict its application [1, 8, 9]. Generally, a surface hardening process is applied to the base metal to eliminate these drawbacks [9, 10]. Additive manufacturing (AM) is currently an attractive production method due to its ability to obtain complex shapes and customized components, which may not be possible with traditional methods [11, 12]. Selective laser melting (SLM) is an additive manufacturing process in which the part is obtained via layer-by-layer processing of the powders utilizing the CAD geometry [13, 14]. The studies in the literature regarding producing 316L stainless steel parts with the SLM process showed that they have better strength and ductility properties than traditionally produced counterparts [15-20]. It has also been reported that tribological properties, i.e., wear resistance, of the 316L manufactured by SLM are superior to cast or hot-sintered products [18]. Considering the abovementioned low tribological characteristics of 316L, parts produced with SLM become advantageous. However, the improvement of the tribological properties of 316L with the help of surface treatment was higher [21, 22] compared to the additively manufactured part [18]. There are various engineering methods for surface hardening of steels, such as thermochemical diffusion, applied energy, coating, and surface modification methods [23]. Boriding is one of the diffusion methods that provides superior tribological properties compared to other conventional thermochemical processes such as carburizing and nitriding [24-27]. The applications of boriding in industry range from industrial machinery to biomedical devices due to frequent exposure to environments such as wear and corrosion. Therefore, it is vital to understand how the boriding process takes place in ferrous or non-ferrous materials used in these areas [28]. Different boriding techniques include gas boriding, molten salt boriding, and solid-state boriding [29]. Pack boriding is a solid-state boriding process in which the sample is embedded in a powder pack containing boron-yielding compositions [30, 31]. The typical process temperature and time ranges are 700-1000 ℃ and 1-12 h, respectively [32]. Additionally, GARCIA-LEÓN et al [33] applied pack boriding on 316L SS at 1000 ℃ for 4 h and showed that the boride layer (FeB+Fe2B) improved the wear resistance by 41 times in comparison with the unborided sample. ARTEAGA-HERNANDEZ et al [6] borided 316L SS with pack boriding at 850, 950 and 1050 ℃ for 2 h, 4 h and 6 h, respectively. It was concluded that borided samples provided better wear resistance, and the improvement became better at low temperatures for longer exposure time or at high temperatures for shorter exposure time [6]. KAYALI et al [34] investigated the corrosion of the borided 316L SS where there was no improvement in the corrosion resistance in a simulation body fluid environment. The kinetics of the boriding process of commercial 316L SS was studied by OZDEMIR et al [35] using boriding conditions of 800, 875 and 950 ℃ for 2 h, 4 h, 8 h, and an empirical equation was proposed to estimate the coating thickness at different temperatures and time. In another study, CAMPOS-SILVA et al [36] investigated the boriding kinetics of growing surface layers during the pack boriding of commercial 316L SS by considering the diffusion zone. IPEK AYVAZ et al [37] concluded that applying the boriding process under microwave heating conditions accelerates the kinetics of the pack boriding of commercial 316L SS, which gives thicker boride layers. The effect of pulsed direct current (DC) during pack boriding of commercial 316L SS was investigated, and it was shown that the growth rate of the boriding layers was improved [38]. In addition to the boriding process of 316L SS samples, the kinetic of layer growth is another important factor that has been worked extensively to predict boride layer thickness in advance. In the literature, there are many studies and reports about the kinetic growth of boride layers on steel substrates [39, 40]. For instance, ORTIZ-DOMÍNGUEZ et al [41] applied two mathematical models to calculate the boride layer thickness and activation energy for ASTM A36 steel. The activation energy was estimated being as 161 kJ/mol. SEN et al [42] studied the kinetic of layer growth on the AISI 4140 steel at 850, 900, 950 ℃ for 2, 4, 6 and 8 h, respectively. The value of activation energy for boriding process of the AISI 4140 steel was determined as 215 kJ/mol. 316L, as mentioned above, has prevalent applications. With AM, it is possible to produce specific parts for the desired applications. However, AM results in different properties compared to traditional manufacturing methods such as microstructure and mechanical properties. Just like these differences, it is important to examine how 316L produced by additive manufacturing will behave against the boriding process to increase or expand the efficiency of the material in the application area. In this context, the study of how the boriding kinetics are realized will open a horizon. During the AM process, the solidification of molten liquids occurs so quickly than casting method. Therefore, fine columnar, molten pool boundaries’ microstructure might play vital role in boriding process of SLM 316L SS.
Within this framework, to the best of our knowledge, there are limited studies regarding the kinetic of boride layer growth for additively manufactured 316L SS. The present study aimed to investigate the kinetic of boride layer growth for SLM 316L SS. Optical microscope, X-ray diffraction (XRD) and field emission scanning electron microscope (FESEM) analysis were performed to reveal the boride layers. Moreover, the kinetic of boride layer growth was also reported.
2 Experimental
2.1 Selective laser melting and boriding
This study used 316L stainless steel powder to produce a rectangular prism sample (13 mm×9 mm×9 mm) with the SLM process, which was performed utilizing a Concept Laser MLAb R machine (trade name CL 20 ES) [43]. The average powder size for 316L SS alloy was 25 µm, ranging from 15 to 45 μm. The chemical composition of the 316L SS powder was given according to the supplier’s data (GE Additive) in Table 1. Also, the average chemical composition of 316L SS powder was obtained by inductively coupled plasma optical emission spectrometry (ICP-OES, Thermo Scientific iCAP 600 Series).
Material | Cr | Ni | Mo | Mn | Si | P | C | S | N | O | Fe |
---|---|---|---|---|---|---|---|---|---|---|---|
Supplier’s data | 16.5-18.5 | 10-13 | 2-2.5 | 0-2 | 0-1 | 0-0.045 | 0-0.03 | 0-0.03 | — | — | Balance |
Powder (ICP-OES) | 17.4 | 12.3 | 2.2 | 1.6 | 0.05 | 0-0.002 | 0.019 | 0-0.001 | 0.06 | 0.045 | Balance |
The laser power was 40 W, and the laser speed was 400 mm/s. The layer thickness and the scan rotation in the process were 25 µm and 90°, respectively. The schematic illustration of the printed sample was shown in Figure 1(a). After 316L SS was printed, the samples to be borided were obtained by cutting pieces from the as-built 316L cuboid. A schematic representation was given in Figure 1(b). The size of the borided samples was 9.0 mm×4.5 mm×2.5 mm. Before the boriding process, each rectangular sample was mechanically polished by SiC grinding paper using 180, 240, 320, 400, 600, 800 and 1000 grit for each surface, respectively and cleaned in an ultrasonic bath using alcohol for five minutes. Then, all of the polished and cleaned samples were dried. The surface roughness, Ra, of the sample was estimated by Ambios XP-2 surface profilometer as about 0.25 µm, which is in agreement with the literature [44]. The pack boriding was performed in a sealed stainless steel (304L) container filled with Ekabor-II powder, and the samples were embedded. The top layer of Ekabor-II powder was also covered with SiC powder to prevent oxidation. The boriding process was carried out in an electrical-resistant furnace at atmospheric pressure, and the process parameters were set up to be 850 ℃ (2, 4 and 6 h), 900 ℃ (2, 4 and 6 h), and 950 ℃ (2, 4 and 6 h). The container was cooled to room temperature in the furnace at the end of the process, and the samples were taken out.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F001.jpg)
2.2 Physico-chemical characterization
The cross sections of the as-built SLM 316L and the borided samples were prepared metallographically by grinding with SiC papers (320, 400, 600, 800, 1000 grit) and final polishing (alumina slurry). The etching was applied using Aqua regia (V(HCl):V(HNO3)=3:1) solution. The microstructural observation was performed with an optical microscope (Olympus) and scanning electron microscopy (SEM) via Carl Zeiss 300 VP at 15 kV accelerating voltage. The energy dispersive spectroscopy (EDS) was also utilized for the elemental analysis. The thickness of the boride layer (FeB+Fe2B) was measured using Image-J software by following the procedure given in the literature [36], and the results were an average of 30 measurements. X-ray diffraction (XRD) analysis was carried out utilizing Rigaku ULTIMA 3-Rint 2200/PC instrument to characterize the as-built and borided samples at 850, 900, 950 ℃ for 4 h. For the XRD measurement Cu-Kα radiation (1.54 Å) at 45 kV-40 mA over a range from 10° to 90° was used. The hardness of the boride layers and substrate was measured on the cross-sections by using microhardness tester (FutureTech FM-ARS 7000). The applied load was 50 g.
3 Results and discussion
3.1 Microstructure of SLM 316L SS
The optical microscope (OM) and SEM images of the as-built SLM 316L SS were depicted in Figure 2. The microstructure images in Figures 2(a) and (b) corresponded to the xz plane, which was parallel to the building direction. SEM image of the as-built SLM 316L SS was shown in Figures 2(c)-(f). As seen in Figure 2(a), it can be observed that the microstructure is composed of columnar grains that exhibit a molten pool structure with semi-elliptical-shaped borders. This kind of microstructure is typical for the AM 316L SS [45, 46]. These molten pools formed and extended along the building direction owing to the temperature gradient.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F002.jpg)
Moreover, as seen in Figure 2(b), fine columnar microstructures in the regions close to the molten pool boundaries occurred, whereas cellular structures formed inside molten pools at high magnification images. In other words, the finer microstructure could be observed near the edges of the molten pool and coarser in the center. The formation of fine columnar microstructures near molten pool boundaries could be ascribed to a rapid cooling rate compared to the upper side of columnar grains [47, 48]. As seen in Figure 2(c), the molten pool boundaries, cellular structures, and columnar grains could be seen as a typical SEM microstructure of AM 316L SS [49, 50]. In addition, Figure 2(d) showed the magnified picture of the area inside the square in Figure 2(c). As seen in Figure 2(d), the columnar structures formed and extended perpendicular to the boundaries of the molten pool. The cellular type structure could be easily observed as well. Furthermore, some of the fine and cellular structures formed along with the molten pool boundary. Figure 2(e) showed the magnified picture of the area inside the square in Figure 2(d). The image at lower magnification was shown in Figure 2(f). According to this image, the sizes of the cellular structures were measured, and their diameters ranged between 750 and 900 μm.
In addition to SEM images of SLM 316L SS samples, the EDS analysis was also conducted, as demonstrated in Figure 3. The EDS spot analysis was performed on the SLM 316L SS samples as seen in the SEM image in Figure 3(a). The EDS analysis spots 1 and 3 were taken from the interior of the cellular structures, while EDS analysis spots 2 and 4 were taken from the border. Figures 3(b)-(g) exhibited the elemental composition obtained from EDS analysis. As seen in Figures 3(b)-(g), the border of the cellular structures was slightly enriched in Cr, Ni, Mo, Mn, and Si.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F003.jpg)
3.2 X-ray diffraction
XRD analysis of SLM 316L SS was demonstrated in Figure 4. As seen in Figure 4, the as-built SLM 316L SS had three major peaks at near 43.44°, 50.57° and 74.49° which corresponded to γ-austenite phase [51]. After the boriding process at determined temperatures (850, 900 and 950 ℃) for 4 h, it can be observed that both FeB and Fe2B layers formed on the surface of the as-built SLM 316L SS. In addition, the formation of CrB and Cr2B borides phases was also detected. As can be seen in Figure 4, the characteristic peaks of Cr2B and CrB generally overlap the peak position of iron borides [34, 52, 53]. Moreover, the intensity of FeB peaks increased as the boriding temperature increased. Moreover, the appearance of B4C compound might be related to boriding environment. Our results were in good agreement with the literature studies [54, 55]. The presence of the phase was also approved by Rietveld analysis as seen in Figure S1. The application of the Rietveld analysis assisted to define the phases formed during boriding process. The existence of both FeB and Fe2B phases indicated the formation of double layers owing to the diffusion of boron atoms into the material. Furthermore, the residual stress analysis of the borided samples was carried out by Rietveld refinement analysis based on XRD results. The micro strain values of the FeB layers were found to be 6.6×10-4, 2.1×10-3 and 3×10-3 for the samples borided at 850, 900 and 950 ℃, respectively for 4 h. It can be concluded that the strain increased as temperature soared. This increment of strain between the layer might lead to high stress and should cause the peeling off or defragmentation of FeB layer as seen in SEM images.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F004.jpg)
3.3 SEM analysis of borided SLM 316L SS
The cross-section SEM images of the borided SLM 316L samples were illustrated in Figure 5. As seen in Figure 5, the morphology of the boride layer-transition matrix interface was smooth and flat because of the alloying element effect [38]. The printed samples generally consisted of 4 different zones. The first zone corresponded to the formation of the FeB layer, represented by the red line, while the second region, represented by the orange line, belonged to the Fe2B layer. It was seen that the FeB and Fe2B layers were in different contrasts. The third region, represented by the green line, and the fourth zone were associated with the transition matrix layer and substrate, respectively. It could be observed that the thickness of each boride layer increased as the boriding temperature and time increased up. There are theoretically two factors in growth reaction kinetics. These are reaction products diffused on the reaction interface (diffusion-controlled) and a chemical barrier for transportation of the substances reacting on the reaction interface (reaction controlled). Boride layer grows linearly as the treatment time increases in the case of reaction-controlled process; it grows proportional to the square root of the reaction time if the process is diffusional. The changes in the thickness of both FeB and Fe2B layers versus boriding time to temperature were given in Figure 6. As seen in Figure 6, the thickness of the iron boride layers grew proportional to the square root of the reaction time. Generally, the thickness of the samples ranged from (1.8±0.3) μm to (27.7±2.2) μm. Furthermore, the defragmentation of the formed FeB layers could be observed depending on the temperature and duration of the boriding process. As the boriding temperature and time increased, the defragmentation of the FeB layer was very severe because of intrinsic properties [56]. The formation of the Fe2B phase is more desirable than the formation of the FeB and Fe2B bi-phase layer on the surface. Because the FeB phase is rich in boron, this phase contains 16.23% boron by weight. This is undesirable because the FeB phase is more brittle than other iron-boron phases [57]. In addition to that, the high temperature and long duration could cause residual stress between the formed boride layers. Therefore, residual stress might lead to fracturing and peeling off the FeB layers at high temperatures and long duration. Similar layer regions can be supported by previous studies [22, 35, 58]. The change of Vickers microhardness profile of a representative sample (900 ℃ for 4 h) versus distance from the surface was shown in Figure S2. While the hardness value of boride layer was obtained to be HV0.05 1900 nearby surface. As it moved away from the surface, a noticeable decrease in the hardness value was observed and the hardness value was determined as 237 HV0.05. This is equivalent to the hardness of the substrate sample. This trend showed that the boride layers were made of different diffusion zones as seen in Figure 5.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F005.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F006.jpg)
In addition to SEM images of borided samples, the EDS analysis was also conducted to prove the existence of the formation of FeB and Fe2B boride layers on the SLM 316L SS alloys in Figure 7. The sample borided at 850 ℃ for 6 h was chosen as an example image to demonstrate boride layers. The amount of boron element for both FeB and Fe2B layers was 17.44 wt.% and 7.06 wt.%, respectively. These are very close to both theoretical weight ratios of FeB and Fe2B structure, which indicated that the dual boride layers occurred on the surfaces.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F007.jpg)
The thickness of the boride layer was measured based on the boriding time at different temperatures. This relationship was illustrated in Figure 6. The values of the boride layers were also provided in Figure 6. It could be said that the initial increase in the thickness of each layer was rapid, but as time progressed, the rate of increase tended to decline. This meant that the thickness of the boride layer changed parabolic with time at all temperatures, as seen in Figure 6.
3.4 Growth kinetics of boride layers on as-built SLM 316L SS
Boriding is a diffusional process. The migration of boron atoms into the material to form boride phases on the surfaces is controlled by the diffusion mechanism in the boriding process [59]. The thickness of the boride layer strongly depends on the boriding medium, the chemical composition of the substrate, temperature, and time [60]. The boron concentration profiles and diffusion zone in FeB/Fe2B bilayers were illustrated in Figure 8. In Figure 8,
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F008.jpg)
Some assumptions are generally taken into account to simplify the kinetic model application. Some of these assumptions are listed as follows [60, 63]: 1) The boron atom’s diffusion governs the boride layers’ growth kinetic into FeB/Fe2B phases; 2) The boron concentration profile is assumed to be linear throughout the FeB and Fe2B layers; 3) The diffusion of boron atoms and the growth of boride layers are perpendicular to the sample surface; 4) The boride layer is thin compared to the sample thickness; 5) The diffusion coefficient of boron in each iron boride is independent of concentration, and the effect of Fe diffusion on layer growth is neglected; 6) The nucleation of iron borides starts after incubation time, and the incubation time is neglected; 7) The temperature is constant and uniform throughout the sample. The growth kinetics of the boride layers into FeB/Fe2B phases is generally controlled by diffusion of boron element as mentioned above. Therefore, Fick’s second law has been used to predict quantitative growth kinetics of boride layers such as FeB and Fe2B layers formed on the surface of Fe substrates. Fick’s second law defines the alterations in the concentration of atoms for time and diffusion position during the boriding process in the following equation [64, 65].
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M006.jpg)
Equation (1) transforms to the general expression after solving it as given in the following equation [66].
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M007.jpg)
where Ci(x, t) expresses the concentration at a certain thickness (x) depending on time (t) and temperature (T). Di describes the diffusion coefficient. Ai and Bi are constants.
According to Wagner’s approach [67], the thickness of FeB and Fe2B layers can be expressed in the following equations concerning time (t).
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M008.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M009.jpg)
where ξ and η are constants, and D1 and D2 are the boron diffusion coefficients in the FeB and Fe2B layers, respectively. D1 and D2 could be expressed in the following equations.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M010.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M011.jpg)
where A1 and A2 are constants; Q1 and Q2 define the activation energy.
The curves of x2FeB versus t and x2Fe2B versus t should be plotted to assess the ξ and η values for the boriding process. For this purpose, the square of boride layer thickness vs boriding time curves were plotted and shown in Figure 9. As seen in Figures 9(a) and (b), the slopes of the straight lines of (x2FeB-t) and (x2Fe2B-t) are equal to 4ξ2D1 and 4η2D2. While the 4ξ2D1 values of FeB layers for 850, 900 and 950 ℃were found to be 7.8×10-16, 4.33×10-15 and 7.243×10-15 m2/s, respectively; the 4η2D2 values were found to 1.47×10-15, 3.6×10-15 and 6.04×10-15 for the Fe2B layer, respectively.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F009.jpg)
The initial and boundary conditions of the boron diffusion can be expressed to determine ξ and η values. The initial conditions of the diffusion problems were set up as follows:
Initial condition: (t=0), for x>0
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M012.jpg)
The following equations also gave the boundary conditions.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M013.jpg)
The two parameters
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M016.jpg)
Equations (2)-(4) were incorporated with the initial and boundary conditions in Eqs. (7)-(9). As a result, the variation in B concentration in the FeB and Fe2B phases can be designated as:
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M017.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M018.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M019.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M020.jpg)
The growth rate of the borided FeB and Fe2B layers could be expressed as follows due to the mass balance equation at the (FeB/Fe2B) and Fe2B/substrate) interfaces [68].
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M021.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M022.jpg)
While x=u, taking the derivation of Eqs. (10) and (12) respectively, Eqs. (15) and (16) were obtained.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M023.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M024.jpg)
While x=v, taking the derivation of Eq. (12), we obtained the following equation.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M025.jpg)
Replacing Eqs. (15)-(17) into Eqs. (13) and (14), and reorganizing give:
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M026.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M027.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M028.jpg)
The ξ and η constants in the Eqs. (18) and (19) were solved numerically using a written computer program in MATLAB software. The values of the ξ and η constants were 0.228 and 2.739, respectively. The diffusion coefficient of boron element (
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M033.jpg)
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M034.jpg)
An empirical equation can be obtained to predict the thickness of the boride layers with the help of kinetic diffusion theory. The empirical equation can be generated by using the substitution Eqs. (20) and (21) into Eqs. (3) and (4) after solving unknown values such as ξ, η and D for experimental results. After determining the unknown values, the empirical equation can be formed, as seen in Eq. (22), and can be used to estimate boride thickness according to initial conditions.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-M035.jpg)
Within this framework, the comparison of the layer thickness estimated from the experiment and empirical results was given in Figure 10. The thickness values obtained from experimental results were close to the thickness values generated from empirical results.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F010.jpg)
In Table 2, some literature findings were given to compare the results of this study and the literature. The thickness of the boride layers ranged from (1.8±0.3) μm to (27.7±2.2) μm in this study. For example, FERNÁNDEZ-VALDÉZ et al [22] determined that the thickness of the total iron boride layer changed between approximately 14 and 47 µm for 2 h and 6 h at 900 ℃, respectively, on the surface of 316L SS. CAMPOS-SILVA et al [38] found the thickness of the total iron boride layer to be ~8 μm at 850 ℃ for 0.5 h on the surface of 316L SS. When the thicknesses of the boride layers at the same time and temperature are compared, it could be observed that the thicknesses of the boride layers in the literature are generally larger than this study. The Q is a critical parameter that affects the rate of boride layer growth during boriding processes. It influences the diffusion kinetics of boron or other elements into the substrate material, which, in turn, determines the thickness of the boride layer formed. Lower activation energy promotes faster growth, resulting in thicker layers, while higher activation energy slows down the growth process, leading to thinner layers [49]. As seen in Table 2, the Q values in this study are generally higher than those in the literature. This could be attributed to the microstructure of the SLM 316L SS. The microstructure of SLM 316L SS can significantly influence the diffusion of boron element. The microstructural features that impact diffusion include grain boundaries, porosity, and the presence of defects. In additive manufacturing, the microstructure often consists of columnar, fine, and cellular grains with distinct grain boundaries because of the rapid cooling. These structures possess higher internal stresses, higher dislocation density, higher defects, and a higher number of cellular grain boundaries and cellular structures. Grain boundaries can be diffusion paths for elements, especially at elevated temperatures. Grain boundary diffusion can be faster compared to diffusion within the grains. The presence of grain boundaries can enhance the diffusion of elements like boron and nitrogen in SLM-printed 316L SS, as well as defects, cellular structure, and dislocation density [71, 72]. Even though the microstructure of SLM 316L SS favors the diffusion process, the results contradict the advantages of the microstructure of SLM 316L SS for the diffusion process in this study. This might be due to the segregation of alloying elements such as Cr, Ni, Mo, Mn, and Si in the cell border for SLM 316L SS. As mentioned above, the presence and formation of cellular microstructure can increase the diffusion of elements such as boron. Nevertheless, simultaneously, it accelerates the migration of atoms such as Cr, Ni, and Mo in the SLM 316L SS towards the grain boundaries. It causes intense segregation in the grain boundary regions due to the continuous remelting and solidification of stainless steel in the additive manufacturing method. The formation of boron compounds of the segregation of alloying elements might reduce and hinder the inward diffusion of boron [73-78].
Steel | Temperature range and time | Boriding method | Activation energy for FeB/ (kJ·mol-1) | Activation energy for FeB+Fe2B/ (kJ·mol-1) | Reference |
---|---|---|---|---|---|
316L | 850, 900, 950 °C; 0.5, 1, 1.5, 2.5 h | Pulsed DC pack boriding | 162 | 171 | [38] |
316L | 800, 875, 950 °C; 2, 4, 8h | Pack boriding | ― | 199 | [35] |
316L | 850, 900, 950, 1000 °C; 2, 4, 6, 8, 10 h | Pack boriding | 204 | 198 | [36] |
316L | 700, 750, 800 °C; 3, 5, 7 h | Plasma Paste boriding | ― | 250.8 | [69] |
316L | 800, 850, 900, 950 °C; 2, 4, 6 h | Pack boriding with microwave heating | ― | 244 | [37] |
304L | 850, 900, 950 °C; 2, 4, 6 h | Pack boriding | ― | 222.8 | [70] |
SLM 316L | 850, 900, 950 °C; 2, 4, 6 h | Pack boriding | 256.56 | 209.014 | This work |
4 Conclusions
In conclusion, 316L SS samples were successfully manufactured by a selective laser melting process. The pack boriding method was used to produce boride layers on the surface of the SLM 316L SS. The kinetics of the boride layer growth rate were investigated. The research findings revealed that:
1) According to OM and SEM analysis, the as-built SLM 316L SS microstructure was composed of columnar grains, which exhibited a molten pool structure with semi-elliptical-shaped borders and cellular structures. These molten pools formed and extended along the building direction owing to the temperature gradient. Fine columnar microstructures occurred close to the molten pool boundaries, whereas cellular structures formed inside molten pools at high magnification images.
_Xml/alternativeImage/4ACEB91A-C0D2-4522-BC04-EADD6501EDC4-F011.jpg)
2) FeB and Fe2B layers formed on the as-built SLM 316L SS surface. In addition, the formation of CrB and Cr2B phases was also detected based on XRD results. Moreover, the intensity of FeB peaks increased as the boriding temperature increased. SEM images showed that the boride layer-transition matrix interface morphology was smooth and flat. The thickness of each boride layer increased as the boriding temperature and time rose. The thickness of the FeB layers ranged from (1.8±0.3) μm to (11.6±1.2) μm, whereas the thickness of the total boride layer (FeB+Fe2B) changed from (7.6±0.4) μm to (25.2±2.2) μm with increased time and temperature. The thickness of the boride layers exhibited parabolic character at all temperatures.
3) A diffusion model in Fick’s second law was applied to predict and elucidate the kinetics of the growth rate of FeB and Fe2B phase layers. The activation energy (Q) values of the individual FeB layer, Fe2B layer, and dual FeB+Fe2B layer were found to be 256.56, 161.61 and 209.014 kJ/mol, respectively. A satisfactory coherence was obtained between the calculated and experimental results.
4) The Q values of the boride layers in this study were higher than those in the literature. The segregation of alloying elements in the border of cellular structures could lead to declining and preventing the inward diffusion of boron.
Micro-abrasion wear resistance of borided 316L stainless steel and AISI 1018 steel
[J]. Journal of Materials Engineering and Performance, 2017, 26(11): 5599-5609. DOI: 10.1007/s11665-017-3004-0.Oxidation behavior of 316L austenitic stainless steel in high temperature air with long-term exposure
[J]. Materials Research Express, 2020, 7(6):Enhanced mechanical properties and corrosion resistance of 316L stainless steel by pre-forming a gradient nanostructured surface layer and annealing
[J]. Acta Materialia, 2021, 208: 116773. DOI: 10.1016/j.actamat.2021.116773.Dry sliding wear test on borided AISI 316L stainless steel under ball-on-flat configuration: A statistical analysis
[J]. Tribology International, 2021, 157: 106885. DOI: 10.1016/j.triboint.2021.106885.Multi-pass scratch test behavior of AISI 316L borided steel
[J]. Surface and Coatings Technology, 2016, 307: 491-499. DOI: 10.1016/j.surfcoat.2016.09.017.Study of boriding surface treatment in the tribological behavior of an AISI 316L stainless steel
[J]. Wear, 2021, 477: 203825. DOI: 10.1016/j.wear.2021.203825.High strength and ductility of additively manufactured 316L stainless steel explained
[J]. Metallurgical and Materials Transactions A, 2018, 49(7): 3011-3027. DOI: 10.1007/s11661-018-4607-2.Development of tribological maps on borided AISI 316L stainless steel under ball-on-flat wet sliding conditions
[J]. Tribology International, 2021, 163: 107161. DOI: 10.1016/j.triboint.2021.107161.Surface engineering of austenitic stainless steel
[J]. Surface Engineering, 2002, 18(6): 415-422. DOI: 10.1179/026708402225006268.BONN technique: Tribological properties predictor for plasma nitrided 316L stainless steel
[J]. Materials Today: Proceedings, 2018, 5(6): 14545-14552. DOI: 10.1016/j.matpr.2018.03.044.Metal additive manufacturing: Technology, metallurgy and modelling
[J]. Journal of Manufacturing Processes, 2020, 57: 978-1003. DOI: 10.1016/j.jmapro.2020.07.025.Microstructural control in metal laser powder bed fusion additive manufacturing using laser beam shaping strategy
[J]. Acta Materialia, 2020, 184: 284-305. DOI: 10.1016/j.actamat.2019.11.053.Additive manufacturing of metallic materials: A review
[J]. Journal of Materials Engineering and Performance, 2018, 27(1): 1-13. DOI: 10.1007/s11665-017-2747-y.Selective laser melting of high strength Al-Cu-Mg alloys: Processing, microstructure and mechanical properties
[J]. Materials Science and Engineering A, 2016, 656: 47-54. DOI: 10.1016/j.msea.2015.12.101.Crystallographic-orientation-dependent tensile behaviours of stainless steel 316L fabricated by laser powder bed fusion
[J]. Materials Science and Engineering A, 2019, 766: 138395. DOI: 10.1016/j.msea.2019.138395.Dislocation network in additive manufactured steel breaks strength-ductility trade-off
[J]. Materials Today, 2018, 21(4): 354-361. DOI: 10.1016/j.mattod.2017.11.004.Additively manufactured hierarchical stainless steels with high strength and ductility
[J]. Nature Materials, 2018, 17(1): 63-71. DOI:10.1038/nmat5021.316L stainless steel mechanical and tribological behavior: A comparison between selective laser melting, hot pressing and conventional casting
[J]. Additive Manufacturing, 2017, 16: 81-89. DOI: 10.1016/j.addma.2017.05.007.Microstructure and mechanical properties of stainless steel 316L vertical struts manufactured by laser powder bed fusion process
[J]. Materials Science and Engineering A, 2018, 736: 27-40. DOI:10.1016/j.msea. 2018.08.069.Mechanical behavior of selective laser melted 316L stainless steel
[J]. Materials Science and Engineering A, 2017, 696: 113-121. DOI: 10.1016/j.msea.2017.04.058.Wear maps of borided AISI 316L steel under ball-on-flat dry sliding conditions
[J]. Materials Letters, 2021, 282: 128842. DOI: 10.1016/j.matlet.2020. 128842.Sliding wear analysis in borided AISI 316L steels
[J]. Materials Letters, 2021, 285: 129138. DOI: 10.1016/j.matlet.2020.129138.Boronizing of iron-based alloys
[M]//Encyclopedia of Iron, Steel, and Their Alloys.An investigation on surface properties of treated low carbon and alloyed steels (boriding and carburizing)
[J]. Journal of Materials Processing Technology, 2000, 103(2): 310-317. DOI: 10.1016/S0924-0136(99)00488-4.Comparison of hardness and wear behaviour οf boronized and carburized AISI 8620 steels
[J]. Acta Physica Polonica A, 2009, 116(6): 1029-1032. DOI: 10.12693/aphyspola.116.1029.Abrasive wear behavior of boronized AISI 8620 steel
[J]. Wear, 2009, 266(11, 12): 1106-1112. DOI: 10.1016/j.wear.2009.03.006.Historical review on the boriding process using bibliometric analysis
[J]. Transactions of the Indian Institute of Metals, 2021, 74(3): 541-557. DOI:10. 1007/s12666-020-02174-6.An overview on surface hardening of titanium alloys by diffusion of interstitial atoms
[J]. Diffusion Foundations, 2015, 4: 103-116. DOI: 10.4028/www.scientific.net/df.4.103.Effect of shot blasting on the boriding kinetics of AISI 316L stainless steel
[J]. Protection of Metals and Physical Chemistry of Surfaces, 2023, 59(4): 671-678. DOI: 10.1134/s2070205123700624.Mechanical characterization of the AISI 316L alloy exposed to boriding process
[J]. Dyna, 2020, 87(213): 34-41. DOI: 10.15446/dyna.v87n213.82924.Corrosion and wear behaviors of boronized AISI 316L stainless steel
[J]. Metals and Materials International, 2013, 19(5): 1053-1061. DOI: 10.1007/s12540-013-5019-x.An investigation on boriding kinetics of AISI 316 stainless steel
[J]. Vacuum, 2008, 83(1): 175-179. DOI: 10.1016/j.vacuum.2008.03.026.Formation and kinetics of FeB/Fe2B layers and diffusion zone at the surface of AISI 316 borided steels
[J]. Surface and Coatings Technology, 2010, 205(2): 403-412. DOI: 10.1016/j.surfcoat.2010.06.068.Effect of the microwave heating on diffusion kinetics and mechanical properties of borides in AISI 316L
[J]. Transactions of the Indian Institute of Metals, 2020, 73(10): 2635-2644. DOI: 10.1007/s12666-020-0207 2-x.Pulsed-DC powder-pack boriding: Growth kinetics of boride layers on an AISI 316 L stainless steel and inconel 718 superalloy
[J]. Surface and Coatings Technology, 2021, 421: 127404. DOI: 10.1016/j.surfcoat.2021.127404.Growth kinetics of iron boride layers: Dimensional analysis
[J]. Applied Surface Science, 2006, 252(24): 8662-8667. DOI: 10.1016/j.apsusc.2005.12.002.Kinetics of the boride layers obtained on AISI 1018 steel by considering the amount of matter involved
[J]. Coatings, 2021, 11(2): 259. DOI: 10.3390/coatings11020259.Modeling of the growth kinetics of boride layers in powder-pack borided ASTM A36 steel based on two different approaches
[J]. Advances in Materials Science and Engineering, 2019: 5985617. DOI: 10. 1155/20 19/5985617.The growth kinetics of borides formed on boronized AISI 4140 steel
[J]. Vacuum, 2005, 77(2): 195-202. DOI: 10.1016/j.vacuum.2004.09.005.Investigation of mechanical, tribological and magnetic properties after plasma nitriding of AISI 316L stainless steel produced with different orientations angles by selective laser melting
[J]. Surface and Coatings Technology, 2023, 467: 129676. DOI: 10.1016/j.surfcoat.2023.129676.Comparison of different post processing technologies for SLM generated 316l steel parts
[J]. Rapid Prototyping Journal, 2013, 19(3): 173-179. DOI: 10.1108/13552541311312166.Effect of build geometry on the microstructural development of 316L parts produced by additive manufacturing
[J]. Materials Characterization, 2018, 143: 137-143. DOI: 10.1016/j.matchar.2018.04.040.Microstructural and mechanical evaluation of post-processed SS 316L manufactured by laser-based powder bed fusion
[J]. Journal of Materials Research and Technology, 2021, 12: 210-220. DOI: 10.1016/j.jmrt.2021.02.090.Influence of heat treatment temperature on the microstructural, mechanical, and wear behavior of 316L stainless steel fabricated by laser powder bed additive manufacturing
[J]. The International Journal of Advanced Manufacturing Technology, 2020, 107(5): 1947-1956. DOI: 10.1007/s00170-020-04972-0.A comprehensive study on microstructure and tensile behaviour of a selectively laser melted stainless steel
[J]. Scientific Reports, 2018, 8(1): 7785. DOI:10.1038/s41598-018-26136-7.Investigating growth of iron borides with the formation of monolithic Fe2B layer on AISI 304 stainless steel via cathodic reduction and thermal diffusion-based boriding
[J]. Journal of Materials Engineering and Performance, 2022, 31(4): 3274-3286. DOI: 10.1007/s11665-021-06417-5.Hardened austenite steel with columnar sub-grain structure formed by laser melting
[J]. Materials Science and Engineering A, 2015, 625: 221-229. DOI: 10.1016/j.msea.2014.12.018.Additively manufactured 316L stainless steel with improved corrosion resistance and biological response for biomedical applications
[J]. Additive Manufacturing, 2019, 27: 8-19. DOI: 10.1016/j.addma.2019.02.005.Characterization and boron diffusion kinetics on the surface-hardened layers of royalloy steel
[J]. Coatings, 2023, 13(1): 113. DOI: 10.3390/coatings13010113.Investigation on microstructure and mechanical properties of boron-modified 13Cr5Ni2Mo by powder-pack boriding
[J]. Materials Science and Technology, 2023, 39(18): 3209-3219. DOI: 10.1080/02670836.2023.2245657.Plasma paste boronizing of AISI 8620, 52100 and 440C steels
[J]. Materials and Design, 2011, 32(4): 2380-2386. DOI: 10.1016/j.matdes. 2010.11.031.A numerical-experimental study of wear resistance of FeB/Fe2B systems
[J]. Journal of Materials Engineering and Performance, 2021, 30(2): 839-849. DOI: 10.1007/s11665-020-05368-7.Studies of boriding using possibility to increase the corrosion resistance of cast steel 20GL
[J]. Coatings, 2022, 12(11): 1789. DOI: 10.3390/coatings12111789.Growth of the Fe2B layer on SAE 1020 steel employed a boron source of H3BO3 during the powder-pack boriding method
[J]. Journal of Alloys and Compounds, 2018, 744: 658-666. DOI:10.1016/j.jallcom. 2018.02.118.Effect of diffusion annealing on wear and cohesion behaviours of boronized AISI 1040 steel
[J]. Tribology International, 2023, 184: 108428. DOI: 10.1016/j.triboint.2023.108428.A diffusion model for describing the bilayer growth (FeB/Fe2B) during the iron powder-pack boriding
[J]. Applied Surface Science, 2005, 252(2): 393-399. DOI: 10.1016/j.apsusc.2005.01.016.FeB/Fe2B phase transformation during SPS pack-boriding: Boride layer growth kinetics
[J]. Acta Materialia, 2005, 53(8): 2361-2368. DOI: 10.1016/j.actamat.2005.01.043.Kinetic modelling of powder-pack boronized 4Cr5MoSiV1 steel by two distinct approaches
[J]. Coatings, 2023, 13(6): 1132. DOI: 10.3390/coatings13061132.Analysis of diffusion coefficients of iron monoboride and diiron boride coating formed on the surface of AISI 420 steel by two different models: Experiments and modelling
[J]. Materials, 2023, 16(13): 4801. DOI: 10.3390/ma16134801.Simulation of the growth kinetics of the (FeB/Fe2B) bilayer obtained on a borided stainless steel
[J]. Applied Surface Science, 2011, 257(6): 2004-2010. DOI:10. 1016/j.apsusc.2010.09.043.Accelerated kinetics of surface hardening by diffusion near phase transition temperature: Mechanism of growth of boride layers on titanium
[J]. Acta Materialia, 2011, 59(10): 4216-4228. DOI:10.1016/j.actamat.2011.03.046.Alternative kinetic model of growth of boride layers on steel AISI 316
[J]. Metal Science and Heat Treatment, 2021, 63(7): 430-436. DOI: 10.1007/s11041-021-00707-4.On the anomalous self-diffusion in B.C.C. titanium
[J]. Physica Status Solidi (b), 1987, 144(1): 243-251. DOI: 10.1002/pssb.2221440122.The evaluation of data obtained with diffusion couples of binary single-phase and multiphase systems
[J]. Acta Metallurgica, 1969, 17(2): 99-107. DOI: 10.1016/0001-6160(69)90131-X.Kinetics of growth of superhard boride layers during solid state diffusion of boron into titanium
[J]. Ceramics International, 2012, 38(8): 6795-6805. DOI: 10.1016/j.ceramint. 2012.05.077.Characterization and kinetics of plasma-paste-borided AISI 316 steel
[J]. Materiali in Tehnologije, 2016, 50(2): 263-268. DOI: 10.17222/mit.2015.031.Investigation of the diffusion kinetics of borided stainless steels
[J]. The Physics of Metals and Metallography, 2013, 114(12): 1061-1068. DOI: 10.1134/S0031918X1322002X.Improving the surface properties of additive-manufactured Inconel 625 by plasma nitriding
[J]. Surface and Coatings Technology, 2023, 452: 129130. DOI: 10.1016/j.surfcoat.2022.129130.Use of plasma nitriding to improve the wear and corrosion resistance of 18Ni-300 maraging steel manufactured by selective laser melting
[J]. Scientific Reports, 2021, 11: 3277. DOI: 10.1038/s41598-021-82572-y.About metastable cellular structure in additively manufactured austenitic stainless steels
[J]. Additive Manufacturing, 2021, 38: 101804. DOI: 10.1016/j.addma. 2020.101804.Microstructure and corrosion behavior of 316L stainless steel prepared using different additive manufacturing methods: A comparative study bringing insights into the impact of microstructure on their passivity
[J]. Corrosion Science, 2020, 176: 108914. DOI: 10.1016/j.corsci.2020.108914.The effect of alloying elements in Ti-5Mo-5V-8Cr-3Al alloy on growth kinetics of TiB whiskers in boride layer
[J]. Materials and Design, 2023, 225: 111478. DOI: 10.1016/j.matdes. 2022.111478.Hydrogen permeation behavior and hydrogen-induced defects in 316L stainless steels manufactured by additive manufacturing
[J]. Materials Chemistry and Physics, 2020, 250: 123038. DOI:10. 1016/j.matchemphys.2020.123038.The effects of boriding process on tribological properties and corrosive behavior of a novel high manganese steel
[J]. Journal of Materials Processing Technology, 2022, 300: 117421. DOI: 10.1016/j.jmatprotec. 2021.117421.Diffusion kinetics of borided AISI 52100 and AISI 440C steels
[J]. Vacuum, 2012, 86(10): 1428-1434. DOI: 10.1016/j.vacuum.2012.03.030.DEMIRCI Selim and TÜNÇAY Mehmet Masum declare that they have no conflict of interest relevant to this article’s content.
DEMIRCI Selim, TÜNÇAY Mehmet Masum. Growth kinetics of borided 316L stainless steel obtained by selective laser melting [J]. Journal of Central South University, 2025, 32(2): 332-349. DOI: https://doi.org/10.1007/s11771-024-5733-1.
.选择性激光熔化制备硼化316L不锈钢的生长动力学[J].中南大学学报(英文版),2025,32(2):332-349.