Open Access

Effects of B content on wear and corrosion resistance of laser-cladded Co-based alloy coatings


Cite

Introduction

Titanium alloy, as the kay matetrial of high- end equipment components such as aeroengine, exhibited good high specific strength, excellent corrosion resistance, super-strong fracture toughness and fatigue performance [1, 2]. Because of the low hardness and poor wear of titanium alloys, its use in high-end applications with high reliability requirements have been limited [3, 4]. The hardness, strength, wear resistance, and corrosion resistance of the surface of the metal parts could be improved by changing, the chemical comoosition and microstructure of the surface of the metal parts through surface modification technology [5, 6], thereby improving the service life of the parts. Surface modification techniques generally included surfacing [7], plasma spraying [8], magnetron sputtering [9], laser cladding [10], and other processes. Laser cladding (LC) is widely used in advanced manufacturing for its low production cost, nonequilibrium rapid cooling, low dilution rate, and heat-affected zone as well as the many cladding materials that can be used [1113]. It is especially suitable for amorphous structures because of their nonequilibrium, rapid cooling characteristics.

LC is an environmentally friendly and economical method for Co-based alloy fabrication, with the good formability and high wear resistance of the cladded coatings [1416]. Co-based alloys showed excellent wear resistance, corrosion resistance, and high-temperature oxidation resistance and had been widely used in aerospace, shipbuilding, petrochemical, and other industries. Although it had been reported that Co-based alloy coatings could improve the performances of the matrix [1719], the mechanical properties of laser-remanufacturing parts would deteriorate due to the existence of slag, cracks, and pores [20]. Therefore, it is of great importance to design Co-based alloys with good formability and high wear resistance.

In order to improve the wear resistance of LC alloy under severe working conditions, amorphous content manipulation and hard phases precipitation were effective means of improving wear resistance enhancement for LC alloys [2124]. Zhang et al. [25] showed that an increase of amorphous phase content in the surface area was beneficial to improving the wear resistance and friction reduction performance of the coating. The hardness and wear resistance of the alloy could also be improved by adding B element because of the formation of hard boride and amorphous elements [26]. Nonetheless, excessive B content would lead to the formation of a large number of brittle particles in the alloy that could reduce the wear resistance resulting from fracture and spalling during the wear process [27]. In the literature cited above, B content exhibited a wonderful contribution to the microstructure and wear resistance of the alloy. It should be noted, however, that there are few reports on the synergistic mechanism of B content on the microstructure, wear resistance, and corrosion resistance of TC4 alloy.

In order to study these mechanisms, (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings with different B contents were prepared on the surface of titanium alloy by laser cladding. In the present study, the effects of B content on the microstructure, wear properties, and corrosion resistance of the coatings were also investigated.

Materials and methods of investigation
Materials

TC4 alloy was selected as the substrate; its main chemical composition is shown in Table 1. The surface of TC4 alloy was polished with sandpaper to ensure a smooth, flat surface. Secondly, ultrasonic cleaning was carried out with acetone solution to remove oil stains on the surface, and the alloy was dried. The laser cladding materials selected in this experiment were Co-based alloy powders with different B contents (Co34Fe8Cr29Ni8Si7)100–xBx (x = 0,4,6,8,12,16), and the particle size of the powder was about 200–400 mesh. The marked numbers and nominal components of the samples are shown in Table 2. The morphology of the B4 powder is shown in Figures 1a, and 1b illustrates its structure. Nd: YAG solid-state pulsed laser (RH-700 type Nd: YAG laser, China) was used in this laser cladding experiment. The processing parameters of the laser cladding included an output current of 380A, a scanning speed of 100 mm/min, a frequency of 4 HZ, a pulse width of 8 ms, and a preset powder thickness of 200 μm. The preinstalled powder and substrate were placed in a chamber made of quartz glass, and argon with a purity of 99.99% was sprayed continuously during cladding to prevent oxidation of the coatings during preparation.

Fig. 1.

The morphology and the XRD test results of the CoCrFeNiSiB4 powder: (a) SEM image, (b) XRD test result [42]

Chemical composition of TC4

Element Fe C N H O Al V Si Ti
wt% 0.30 0.10 0.05 0.01 0.20 5.5–6.8 3.5–4.5 0.15 Bal.

The marked numbers and nominal components of the samples

Sample number Powder(wt%)
B0 Co34Fe8Cr29Ni8Si7
B4 (Co34Fe8Cr29Ni8Si7)96B4
B6 (Co34Fe8Cr29Ni8Si7)94B6
B8 (Co34Fe8Cr29Ni8Si7)92B8
B12 (Co34Fe8Cr29Ni8Si7)88B12
B16 (Co34Fe8Cr29Ni8Si7)84B16
Methods

To study the microstructure of the alloy coatings, a wire-cutting machine was used to make the alloyed coatings into a sample with a size of 10 ×10 × 10 mm. The cross-section and surface of the samples were successively polished with a different mesh of metallographic sandpaper, and mechanical polishing was carried out. Aqua regia was used to etch the coating (HNO3: HCl = 1:3) for 90 s. The microstructure of the samples was characterized by scanning electron microscope (SEM, Zeiss-MERLIN Compact, Germany) equipped with an energy dispersion spectrometer. The phase analysis was performed by Cu-Ka radiation on an X-ray diffractometer (XRD, D/max 2500, Japan). The XRD spectra were measured in 2θ range from 10° to 90° with a step size of 0.02° and a scanning speed of 5°/min. Transmission electron microscopy (TEM, JEOL-2100, Japan) was used to verify some phases in the coatings.

The microhardness was measured with a microhardness tester (TH701, China) under an applied load of 200 g and a residence time of 10 s. According to the standard of ASTM D1044-18, the friction and wear tests were carried out on a high-temperature abrasive testing machine (HT-1000, China) at a temperature of 20°C, a load of 1150 g, a rotation speed of 1120 r/min, and a wear time of 40 min. ZrO2 balls with a diameter of 2.5 mm were selected as grinding materials. The electrochemical corrosion test was carried out in the electrochemical workstation (Zahner, Germany) with 3.5 wt% NaCl electrolyte. A three-electrode device was used in the experiment, in which the assistant electrode was a platinum plate, and the reference electrode was a saturated calomel electrode. Before the measurement of the polarization curves, the samples were immersed in the solution for 20 min, followed by the measurement of the open-circuit potentials (OCP) in the solution for 1 h. Potentiodynamic polarization curves were tested at a scanning rate of 1 mV/s in the range of -0.9 V to 0.3 V.

Results and discussion
Microstructure characteristics

Figure 2 shows the XRD spectrum of the designed alloys, indicating that the coatings mainly consisted of TiCr, Fe0.1Ti0.18V0.72, CoTi, Ti2Ni. A phase of TiB would appear with the gradual increase of B content. The reason was that the hard phase TiB was formed in the process of laser cladding due to the decrease of the Co-based alloy element content and the increase of B content. TiB-reinforced titanium matrix composites were favorable for improving the coatings in this study because of their excellent properties [28]. XRD patterns showed that a wide diffraction peak appeared in the 2θ region, ranging from 44° to 45°, which was attributed to the existence of the amorphous phase in the coatings. At the same time, a strong crystal diffraction peak was superimposed on the diffuse peak, indicating the coexistence of amorphous and crystalline phases in the coatings. The volume percentage of amorphous phase in the coatings was calculated by Pseudo-Voigt function fitting with the method of Verdon with the calculation results shown in Table 3. It could be seen that with the increase of B content, the amorphous content in the coatings increased first and then decreased. When the B content was 6%, the volume fraction of amorphous phase in the coating was highest at around 65.86%.

Theoretical amorphous content of specimens

Sample B0 B4 B6 B8 B12 B16
Content (%) 2.46 35.65 68.86 22.31 10.55 0.97

Fig. 2.

XRD spectrum of specimens

Figure 3 shows the microstructure of three different regions of a typical laser coating crosssection. Figure 3 panels b, c, and d are the enlarged SEM images of the B, C, and D regions in Figure 3a, respectively. The coating exhibited excellent metallurgical bonding with the substrate with no pores. Nevertheless, because of the different thermal expansion coefficients of the substrate and coating materials, as well as the high heating and cooling rates of LC, a small number of cracks appeared in the coating [29] as shown in Figure 3a. Numerous dendrites were found in the middle and bottom regions of the coating, while the top region of the coating was mainly composed of equiaxed crystals. According to the theory of metallurgical solidification, the growth of coating microstructure is closely related to the temperature gradient and solidification rate [30]. The components near the heat-affected zone were not undercooling because of the large temperature gradient and low solidification rate. As a result, the interaction between laser and cladding powder could facilitate the flow and agitation of the molten pool; and the microstructure of the coating exhibited columnar morphology, honeycomb morphology, or dendritic morphology. As a result of the high undercooling rate, the microstructure at the top of the coating was mainly equiaxed crystal.

Fig. 3.

SEM image of the B0 coating: (a) cross sectional morphology of the coating, (b) magnification of rectangular region B, (c) magnification of rectangular region C, (d) magnification of rectangular region D

Figure 4 shows the microstructure of Co-based coatings with different B contents. As shown in Figure 4a, the coating without B element exhibited a typical equiaxed crystal structure, and there was no amorphous region. As shown in Figure 4b, with the addition of B element, the dendrite growth in the coating was improved, forming a microstructure in which the amorphous phase and the crystalline phase were nested and wrapped. The B6 coating was very bright and smooth with no existence of dendrite, as shown in Figure 4c. Therefore, it could be concluded that the coating was mainly composed of amorphous phase, or the corrosion resistance of the coating was higher than that of other coatings. From Figure 4d, e, and f, with the further increasing of B contents, the microstructure of the coatings showed dendrite structure again, and the crystallinity was significantly improved. It could be seen that a change of B contents would affect the microstructure of Co-based alloy coatings, which might exert different effects on the coating properties. Figure 5 shows the SEM image and EDS elemental distribution of the B6 coating, indicating that the elements of B6 coating were uniformly distributed in the amorphous region, and there was basically no large-scale segregation of Ti, Co, Fe, Cr, Ni, or Si metal elements. There was a slight enrichment of B element, which further proved that the addition of an appropriate amount of B element could effectively inhibit the segregation and crystallization of metal elements and improve the formation ability of Co-based amorphous phases.

Fig. 4.

Microstructure of the coatings: (a) B0, (b) B4, (c) B6, (d) B8, (e) B12, (f) B16

Fig. 5.

SEM images and EDS elemental distribution maps of B6 coating

In order to confirm that the coating contained amorphous phase, the B6 coating was studied in detail using TEM. Figure 6a was the bright field TEM image from the upper region of the B6 coating, and Figure 6b gives detailed information about the border between the amorphous phase and the crystalline phase in Figure 6a. The white bright spots in Figure 6a were identified as amorphous. The atomic arrangement in the amorphous phase was disordered, so the bright field region was determined as the amorphous region in Figure 6b. In contrast, the atoms in the dark field were arranged in an orderly fashion, which was determined to be the crystalline area. The existence of both crystalline and amorphous phases in the coating was further confirmed by TEM.

Fig. 6.

TEM image of the B6 coating: (a) the bright field TEM image, (b) the enlarged morphology of the amorphous phase and crystalline phase boundary in (a)

Mechanical properties

The microhardness distribution of (Co34Fe8Cr29Ni8Si7)100–x Bx alloy coatings presented a typical step-like pattern that was about 200 μm as shown in Figure 7. The hardness of the Co-based coatings first increased and then decreased with the increase of B content. When the B content was 6%, the hardness of the B6 coating was up to 1210 HV0.2 which was 3.4 times that of the substrate (350 HV0.2). The microhardness of the coatings was significantly improved due to the combination of grain refinement strengthening, solid solution strengthening, and second-phase strengthening by the hard compound TiB. Certainly, the different content of amorphous phase was also one of the reasons why the microhardness of the coatings was unequal. The coating deviated from the nominal composition, and the microhardness decreased significantly because of the mutual diffusion of elements between the coating and the substrate during LC.

Fig. 7.

Cross-section microhardness curves of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings

The friction coefficient of the (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings first increased and then decreased along with the increase of B contents as shown in Figure 8. Because of the peeling and wear of the coating surfaces during the friction process, the friction coefficient curve fluctuated slightly and finally stabilized at about 0.18. The friction coefficient of B6 coating was the highest, and the friction curve had gone through a gradually declining process. However, the wear loss was the decisive parameter to evaluate wear performance. Under the same conditions, the smaller the wear loss, the higher the wear resistance. The wear loss of B6 coating was as low as 5.8 mg which was one-fourth of that of substrate as shown by the wear loss of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings and substrate presented in Figure 9. To further analyze the wear mechanism, the wear surface was observed using SEM.

Fig. 8.

Friction coefficient curves of the (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings

Fig. 9.

Wear loss of the (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings and substrate

Figure 10 shows the wear morphology of the (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings. As shown in Figure 10a, severe plastic deformation, spalling, and cracks along the wear direction were observed on the wear surface of B0 coating, indicating that serious adhesive wear and fatigue wear occurred during the wear test. According to EDS analysis, the oxygen content of the B0 coating was higher than that of the B6 coating, consequently the B0 coating was also accompanied by oxidative wear as shown in Table 4. The brittle Fe0.1Ti0.18V0.72 intermetallic compounds with dendrite structure in the coating would give birth to stress concentration, which in turn led to the initiation of fatigue cracks on the coating surface and finally extended to the surface along the direction of friction, resulting in spalling of the coating. At the same time, under the rolling of the grinding ball, the formation of abrasive particles in the spalling pit intensified the wear, resulting in the increase of the friction coefficient of the coating and relatively high wear loss [31]. Unlike with B6 coating, cracks were not found on the wear surface of B4 coating as shown in Figure 10b. Shallow furrow tracks along the wear direction were observed, however, indicating that the wear mechanism of B4 coating was abrasive wear. Figure 10c shows the wear surface of the B6 coating with few grooves, which was much smoother as compared with the other coatings. As seen in Figure 10 (d, e, and f), the wear morphology of each sample was similar, but the deformation was smaller than that of B0 coating. The curved grooves, fine debris, and scratches on the whole wear surface were observed, indicating that the wear mechanism was both abrasive wear and slight adhesive wear. Due to the addition of excessive B element, the content of amorphous phase in the coatings began to decrease, the density of defects such as dislocations increased, and brittle intermetallics and borides began to form again. The coatings with less hardness would produce higher wear during use, but the wear resistance of the coatings containing boride and a small amount of amorphous phase was still higher than that of the B0 coating.

EDS analysis in Fig. 9 (wt%)

Ti Cr Fe Co Ni O
A 64.41 1.26 2.78 4.89 5.33 21.33
B 67.22 4.30 1.24 5.95 7.61 13.69

Fig. 10.

Wear morphology surface of the coatings: (a) B0, (b) B4, (c) B6, (d) B8, (e) B12, (f) B16

Based on the above test results, the wear morphology of the coatings was closely related to the content of internal amorphous phase. The addition of B element was theoretically beneficial to the formation of amorphous phase and the reduction of micro-defects of Co-based coatings. The coatings exhibited different wear mechanisms which gradually transitioned from severe fatigue spalling and oxidative wear to slight abrasive wear. When the content of B reached 6%, the coating showed the best wear resistance, and the wear mechanism was mainly slight abrasive wear.

Corrosion behavior

Figure 11 shows the polarization curve of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings in 3.5 wt% NaCl solution at room temperature together with the corresponding electrochemical parameters as shown in Table 5. According to the electrochemical corrosion theory, the free corrosion current density (Icorr) of the alloy is the standard for measuring the corrosion rate, and smaller Icorr implied stronger corrosion resistance [32]. A comparison of the corrosion resistance of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings demonstrated that the Icorr value first decreased and then increased with the increase of B contents, indicating that the addition of B could improve the corrosion resistance of Co-based alloy coatings within a certain range. Under the same corrosion conditions, the Icorr of B4 coating was the lowest at 5.846 × 10–5 A/cm2, indicating that the B6 coating had provided excellent corrosion protection effect on the substrate in comparison to other coatings.

The electrochemical parameters of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings in 3.5 wt% NaCl

Sample Ecorr/mV Icorr/(A·cm–2)
B0 –482.034 8.974×10–5
B4 –419.289 5.846×10–5
B6 –380.255 7.096×10–5
B8 –501.733 8.318× 10–5
B12 –449.567 8.851 × 10–5
B16 –436.434 8.337 × 10–5

Fig. 11.

Polarization curves of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings

Discussion

Amorphous metal alloys, or so-called metal glass (MG), were recognized to have excellent properties due to their unique atomic structure, such as high strength, high hardness, and strong corrosion resistance [3337]. Generally, the difficulty of forming amorphous metal alloys was usually represented by glass-formation ability (GFA) [38]. According to Gibbs free energy formula [39], coating materials with higher mixing entropy and lower mixing enthalpy showed better GFA. The study had also shown that high condensation rate and low dilution rate could significantly improve the GFA of the coatings [40]. Therefore, the Co-based coatings with high mixing entropy and low mixing enthalpy prepared by LC technology in this paper completely conformed to the conditions of amorphous formation. Under the same compressive stress, unlike the crystalline materials with dislocation slip and plastic deformation on the macro and micro scales, the coatings with high amorphous content exhibited less deformation and high hardness due to the absence of defects such as dislocations [41]. This was very consistent with the results shown in Figure 7. Secondly, under the influence of matrix melting and dilution, the chemical composition at the bottom of the molten pool clearly deviated from the deep eutectic composition, which reduced GFA and increased the critical cooling rate required for the formation of amorphous alloy. However, the chemical composition in the upper part of the coatings still maintained a relatively pure original composition due to low dilution, so the amorphous phase was mainly concentrated in the region. Overwhelmingly higher amorphous content had given birth to better wear resistance, yet the crack resistance deteriorated, which made the deeper corrosion of Ti alloy substrate and resulted in its relatively lower corrosion resistance.

Conclusions

The laser-cladded coatings consisting of TiCr, Fe0.1Ti0.18V0.72, CoTi, Ti2Ni, TiB, and amorphous phases exhibited excellent metallurgical bonding with the substrate, in which the TiB and amorphous phases gradually appeared with the addition of B element.

The microhardness of the (Co34Fe8Cr29Ni8Si7)100-xBx alloy coatings increased from 711 HV0.2 to 1210 HV0.2 with increasing B content. When the B content was 6%, the microhardness of the coating increased up to 1210 HV0.2 which was 3.4 times of that of TC4 alloy substrate and 1.7 times of that of CoFeCrNiSi alloy coatings.

With increasing B content, the wear resistance of the (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings were enhanced until the content of B element reached 6%. The wear mechanism consisted of slight adhesive increased amorphous content, which resulted in a wear loss of only 5.8 mg, which was one-fourth of that with the TC4 alloy substrate.

With increasing B content, the corrosion resistance of (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings in 3.5 wt% NaCl solution increased at first and then decreased. Since the volume fraction of amorphous phase was highest in the coating with 4% B content, the lowest free corrosion current density was obtained in 3.5 wt% NaCl solution as compared with that of the other coatings. Comprehensive properties of laser-cladded (Co34Fe8Cr29Ni8Si7)100–xBx alloy coatings had been improved with optimization of the B element content.

eISSN:
2083-134X
Language:
English
Publication timeframe:
4 times per year
Journal Subjects:
Materials Sciences, other, Nanomaterials, Functional and Smart Materials, Materials Characterization and Properties