Next Article in Journal
Thickness and Humidity on Proton Conductivity in MOF-508 Thin Film by Twin-Zinc-Source Method
Previous Article in Journal
Comparative Analysis of Performance of Water-Based Coatings Prepared by Two Kinds of Anti-Bacterial Microcapsules and Nano-Silver Solution on the Surface of Andoung Wood
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Properties of Textured Ni–W Coatings Electrodeposited on the Steel Surface from a Pyrophosphate Bath

1
Hunan Institute of Engineering, College of Materials and Chemical Engineering, Xiangtan 411104, China
2
The Third Oil Production Plant of Changqing Oilfield Company, Yinchuan 750006, China
3
Gas Production Plant 2 of Changqing Oilfield Company, Yulin 719000, China
4
Hunan Nanofilm New Material Technology Co., Ltd., Changsha 410200, China
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(9), 1519; https://doi.org/10.3390/coatings13091519
Submission received: 14 June 2023 / Revised: 23 August 2023 / Accepted: 25 August 2023 / Published: 29 August 2023
(This article belongs to the Section Corrosion, Wear and Erosion)

Abstract

:
Ni–W alloys with a (2 2 0) or (1 1 1) preferred orientation growth and amorphous structure were prepared from a pyrophosphate bath using the electrodeposition method. Structure transformation can be the result of the bath temperature (Tb) and the concentration of sodium tungstate (CW) in the bath. Increasing the Tb and CW can change the crystal growth from (2 2 0) to (1 1 1). At a higher Tb and CW, an amorphous Ni–W alloy can be obtained. The tungsten content in the coatings should be responsible for the structure change. The three textured Ni–W alloys with a (2 2 0) texture, (1 1 1) texture and amorphous structure were annealed at different temperatures ranging from 200 to 700 °C. The microhardness, corrosion resistance and HER of the as-deposited and annealed Ni–W alloys were comparatively studied. The results show that the microhardness of the amorphous Ni–W alloy is the highest and reaches 1028 HV after annealing at 400 °C. The (2 2 0)-textured Ni–W alloy has the best corrosion resistance, which is further improved after annealing, while the HER activity of the (1 1 1) textured Ni-W alloy is superior.

1. Introduction

A material’s texture has remarkable effects on its HER [1,2], corrosion [3,4,5], and other properties [5]. The properties of the HER in an alkaline solution are even changed by macroscopic tensile strains in nickel foils [2]. Xu et al. [1] reported that Ni with a (1 1 1) or (2 0 0) plane-preferred orientation was electrodeposited by changing the additives of NH4Cl and (NH4)2SO4. They found that the (111) preferential orientation promoted the electro-oxidation of NaBH4, and the (200) and (220) planes, as preferred orientations, achieved optimal HER activity. After electrodeposition for 16 h, even the preferential orientation of electrodeposited Ni is changed, from the (111) and (200) planes to the (220) plane [4]. In the simple Watts bath, Nasirpouri [5] prepared Ni films with (1 1 1) and (1 0 0) preferential orientations by pulsed (PC) and pulsed reverse (PRC) current techniques, respectively. Eckold et al. [3] reported that the texture of electrodeposited Sn can be changed by increasing the current density, and a preferred orientation along the lattice planes (321) and (220) greatly enhances the corrosion resistance of the tin layer. In general, metals and alloys with different textures can be prepared by changing the additives and applied current density.
Compared with Ni coatings, Ni–W alloy coatings possess better wear resistance, corrosion resistance [6,7], and hydrogen evolution reaction (HER) activity [8,9,10,11]. They are also considered to be candidates to replace the environmentally hazardous hexavalent hard chromium coating [12]. Differently textured Ni–W alloys have been obtained from an ammonium-citrate bath by DC and PC electrodeposition [6,7]. However, in the bath, it is difficult to solve the problems of citrate decomposition, which leads to plating solution failure, decreased coating performance, and environmental pollution. As a non-cyanide complexant, pyrophosphate is not easily decomposed and consequently presents a longer lifespan. Ni–W crack-free coatings have been prepared from pyrophosphate baths [13,14].
Electrodeposited Ni–W alloys can be improved by heat treatment [15,16,17,18] and the introduction of other elements or compounds [17,18,19,20,21,22,23]. The synergistic effects and reinforcing phases of other elements and compounds can greatly improve the properties of Ni–W alloys. Heat treatment can cause changes in the metal microstructure, such as phase transformation, grain size, residual stress and cracks, consequently optimizing the properties of Ni–W alloy coatings by designing heat treatment processes [6].
In this work, a pyrophosphate bath was selected to prepare Ni–W alloys, and changing the concentration of sodium tungstate (CW) and the bath temperature (Tb) could tune the structure of the Ni–W alloys, consequently forming a preferred orientation of the crystal face, e.g., (2 2 0), (1 1 1), or an amorphous structure. Heat treatment of these Ni–W alloys was carried out at 200 to 700 °C for 2 h in air. The microhardness, corrosion resistance, and HER of the as-deposited and annealed Ni–W alloys were comparatively studied.

2. Experimental Procedure

2.1. Preparation of Ni-W Alloy Coatings

The Ni–W coatings were electrodeposited from a pyrophosphate bath onto Q235 carbon steel disks with a diameter of 10 mm and a thickness of 2 mm. The chemical components of the Q235 steel are listed in Table 1.
Compared with the conventional ammonium-citrate system, the merit of the pyrophosphate bath is that there are no electro-productions resulting from the decomposition of organic chelating agents. To obtain different textured Ni–W alloys, a high bath temperature of 60 °C was also applied in this work. The hydrolysis of pyrophosphate is likely to occur at over 60 °C, so higher temperatures were not selected. The additive is only 2-butyne-1, 4-diol at 0.2 g L−1. The bath composition and the electroplating parameters are listed in Table 2.
Before electrodeposition, the carbon steel disks were ground using SiC waterproof sandpaper (CW-800), electrochemically degreased in 40 g L−1 NaOH solution at a current of 10 A dm−2 for 10 min, and then activated in 10% HCl solution for 10 s. Ni–W alloys were electrodeposited onto the pretreated carbon steel disks in a 500 mL pyrophosphate bath using a DC power supply (APS3005DM, ATTEN/Antaixin, Hangzhou Hamat Electronics Co., Ltd., Zhejiang, China). The bath temperature (Tb) was maintained at 40 or 60 °C. The electrodeposition time was 150 min to meet the requirements of heat treatment and microhardness testing. After electrodeposition, part of the carbon steel disks coated with Ni–W alloys were annealed at 200, 300, 400, 500, 600 and 700 °C for 2 h in a muffle furnace and then placed in a self-sealing bag for storage.

2.2. Characterization of Ni–W Coatings

The chemical composition of the Ni–W coatings was analyzed using an energy-dispersive X-ray spectrometer (EDS, EDX-8000, Shimadzu, Shanghai, China) in an air atmosphere with a collimator of 3 mm. The crystal structure was characterized by X-ray diffraction (XRD, D8 Advance, Bruker, Shanghai, China) in the 2θ range of 20 to 90° (Cu Kα radiation, λ = 0.15406 nm), and the grain size of the strongest diffraction peak was calculated by means of the Debye–Scherrer formula (D = 0.89 λ/ (β cosθ). The microscopic morphology was observed by field emission scanning electron microscope (FESEM, Nova450, FEI). The microhardness of the coatings was tested using a microhardness tester (HV-1000A, Laizhou Huayin Testing Instrument Co., Ltd., Laizhou, China) under an indentation load of 100 gf and a dwelling time of 10 at five different locations per specimen, and the average value was quoted as the hardness of the coating. The samples annealed above 400 °C were polished with W220, W7 and W3.5 metallographic sandpapers in turn because of surface oxidation. Thermo-analysis tests were carried out using a thermal analyzer (STA449F3, Netzsch, Shanghai, China) in air from room temperature to 800 °C at 10 °C min−1. In order to carry out the thermo-analysis tests, the Ni–W alloys were electrodeposited onto stainless-steel disk and were then peeled from the substrate. The weight of all deposits for the thermo-analysis tests was about 10 mg.
The Q235 carbon steel disks coated with Ni–W alloy coatings were sealed with epoxy resin after welding with copper wire and polished with SiC waterproof sandpapers (CW-600, 800, 1000) to expose a sample surface of 0.785 cm2. The electrochemical performance of the sample was tested using an electrochemical workstation (CHI 660E, Shanghai Chenhua, Shanghai, China). The saturated calomel electrode (SCE), platinum plate, and Fe coated with a Ni–W alloy electrode were directly used as a reference electrode (RE), counter electrode (CE) and working electrode (WE), respectively. The corrosion resistance of the Ni–W alloys was evaluated by polarization curves (Tafel) and electrochemical impedance spectroscopy (EIS) in a 3.5 wt. % NaCl solution. The Tafel was tested at a scan rate of 1 mV s−1 with a relative open circuit potential (OCP) of ±250 mV, and the EIS was used with an AC signal amplitude of 5 mV at a frequency range of 0.01 Hz to 100 kHz. The behavior of HER was evaluated by linear sweep voltammetry (LSV) at a scan rate of 5 mV s−1 in a 6 wt. % NaOH solution.

3. Results and Discussion

3.1. Structure of Three-Textured Ni–W Alloy Coatings

3.1.1. As-Deposited Ni–W Alloy Coatings

Figure 1 shows the XRD patterns of the Ni–W alloy coatings prepared at different current densities (4, 6, 8 and 10 A dm−2) and different concentrations of sodium tungstate (CW) (30 or 60 g L−1) at a Tb of 40 or 60 °C. It can be seen that the Ni–W coatings present diffraction peaks at 2θ of about 44°, 51°or 76°, corresponding to the three standard diffraction peaks of pure FCC nickel (JCPDS No: 04-0850). But, the peak position shifts towards a low angle, suggesting FCC lattice shrinkage because of the introduction of large atoms into the Ni matrix. Meanwhile, the higher diffraction angle compared to that of Ni17W3 (JCPDS No: 65-4828) implied a lower W content in the electrodeposited Ni–W alloys.
At a low CW of 30 g L−1 and a low Tb of 40 °C, a (2 2 0) textured Ni–W alloy was obtained, presenting an XRD peak at 2θ = 76°, as shown in Figure 1a. However, the Ni–W alloys obtained when CW = 20 g L−1 and Tb = 40 °C in Ref. [14] present the preferred (1 1 1) orientation. Compared with both baths, the difference is the applied additive, which is 2-butyne-1, 4-diol in this work and is a commercial additive for Ni–W alloys. This result indicates that the additive has an important effect on the growth of Ni–W alloy crystals.
When increasing the CW from 30 to 60 g L−1, the texture of a Ni–W alloy converts to the (1 1 1) orientation, as shown in Figure 1b. The intensity of the (1 1 1) peak becomes strong, and there is an increase in current density, implying the better crystallinity of the Ni–W alloy. However, although the corresponding Ni–W alloys reported in Ref. [14] exhibit the preferred (1 1 1) orientation, the peak intensity decreases with the increase in current density. The difference could result from the different additive. In the present system, the textured transformation from (2 2 0) to (1 1 1) should be attributed to the improved W content in the Ni–W alloys.
When increasing the Tb from 40 to 60 °C, the texture transformation from (2 2 0) to (1 1 1) can also achieve success, as shown in Figure 1c. When increasing the current density to 10 A dm−2, a slight peak of (2 0 0) and a strong peak of (2 2 0) appear. It can be seen that the (1 1 1) peak slightly shifts towards a high angle, along with an increase in current density, suggesting lower W content in the Ni–W alloys.
When simultaneously increasing the CW and Tb, the (1 1 1) peak becomes wide and low, as shown in Figure 1d, implying the formation of an amorphous structure. The higher CW and Tb is beneficial to the W deposition during the electrodeposition of the Ni–W alloy. Under this condition, the (1 1 1) peak hardly rises with the current density; that is, current density has no effect on the crystallinity of Ni–W alloys.
In general, it can be considered that the structure change is mainly attributed to the increase in W in the Ni–W alloys or the selected additive.
The chemical composition of the Ni–W alloy coatings was measured by an energy-dispersive X-ray spectrometer in an air atmosphere with a collimator of 3 mm. The W and Ni content was calculated according to the characteristic peaks at 7.47 KeV for Ni and 9.67 KeV for W. The results are shown in Figure 2. At a low CW of 30 g L−1, as shown in Figure 2a, there is a decreasing trend for the W content along with an increase in the current density, while the trend is weak and even reversed at the high CW of 60 g L−1 (see Figure 2b). At the low CW of 30 g L−1, the W content increases along with the increase in Tb under the same current density conditions, as shown in Figure 2a. However, when increasing CW to 60 g L−1, it is difficult to increase the W content in Ni–W alloys by increasing Tb, especially over 50 °C. These results are consistent with the above XRD data. The diffusion control of the reduction in WO42− ions and the limited deposition of tungsten induced by nickel is a reasonable explanation. Increasing the current density, the surface CW becomes low because of the slow diffusion rate of species containing WO42− ions, consequently resulting in the low W content in Ni–W alloys. When increasing the Tb, the surface CW becomes high due to the increased diffusion rate from the Tb; therefore, the W content in the Ni–W alloys increases. At a high CW of 60 g L−1, the surface CW is high and the deposition of tungsten is limited by nickel, so the W content in the Ni–W alloys is no longer changed by the Tb.
By controlling the CW and the Tb, three textured Ni–W alloys could be obtained and were named W220, W111 and WAS with a (2 2 0) texture, (1 1 1) texture and amorphous structure, respectively. The electrodeposition parameters and the W content in the Ni–W alloys are listed in Table 3.
Figure 3 shows the SEM images of the three textured Ni–W alloy coatings. It can be seen from Figure 3a,b that crystal particles with sizes between 20 and 200 nm gathered together. However, many micron particles can be observed from the surface of the Ni–W alloy coating under the without-additives condition [14]. The result indicates that the crystal size of the Ni–W alloys can be markedly refined by the additives of 2-butyne-1, 4-diol. But, for the WAS sample seen in Figure 3c, the crystal grains cannot be observed from the image with the same magnification (50,000×), and the surface is compact and flat, being related to the amorphous structure. As shown in Figure 3d, the thickness is estimated to be 66, 68 and 81 μm for the W220, W111 and WAS samples, respectively, judging by the cross-sectional view.

3.1.2. Annealed Ni–W Alloy Coatings

Figure 4 shows the XRD pattern evolution of the W220, W111 and WAS Ni–W alloys, along with the increase in annealed temperature. From Figure 4a, it is found that below 700 °C, the (2 2 0) preferred orientation became stronger and stronger when increasing the annealed temperature, while other diffraction peaks became slightly stronger. During heat treatment, the (2 2 0) crystal plane grew rapidly, and the (2 2 0) textured structure could be retained until a temperature of 600 °C was reached. At 700 °C, the Ni–W alloy was oxidized, and new diffraction peaks appear in the magnified curve [24].
As shown in Figure 4b, there is the (111) preferred texture in the W111 sample until a temperature of 500 °C is reached. During heat treatment below 500 °C, there is almost no growth of the (111) crystal surface, while the (2 0 0) and (2 2 0) plane distinguishably grew. The oxidization of Ni–W also happened at 700 °C.
For the WAS sample, the broad and weak (1 1 1) peak becomes strong and narrow with an increasing annealed temperature, suggesting that fine grains were formed during the transformation from the amorphous to the nanocrystalline structure. The (2 0 0) and (2 2 0) peaks are weaker than those of the W111 sample, implying that their growth is suppressed. At 600 °C, the strongest (1 1 1) peak suggests that the transformation from amorphous to crystalline was completed.
The differential scanning calorimetry (DSC) and thermogravimetry (TG) curves of the W220, W111 and WAS samples are shown in Figure 5. It can be seen from the DSC curves that all samples present an exothermic process, which accelerates at about 500 °C. The exothermic process is related to crystal growth, so the preferential diffraction peaks become strong and are strongest after 500 °C, as shown in Figure 4. The severe oxidation at 700 °C made the surface of the Ni–W alloys black (see the inset in Figure 5) and undermined the diffraction peaks of the Ni–W alloys. As a result, the strongest diffraction peaks appear at 600 °C, as shown in Figure 4.
The TG curves present the different shapes. It can be seen from Figure 5 that for the W220 sample, there is almost no change in weight below 335 °C, and then, the weight increases. At the same time, the surface of W220 appears red and even brown at 400 °C, as shown in the insets of Figure 5a.
For the W111 and WAS samples, the weigh firstly decreases, plateaus, and then increases along with the temperature, as shown in Figure 5b,c. The weight loss may result from airborne desorption because the (111) crystal face possesses higher activation than the (220) crystal face. For the WAS sample, more weight loss is observed, and the formation of volatile tungsten oxide on the surface is a reasonable explanation. The initial stage of the plateau corresponds to the start of oxidation because of the appearance of surface color at about 400 °C. At 400 °C, the color becomes gradually light for the W220, W111 and WAS samples, which is consistent with the increase in W in the alloys. At over 600 °C, the rapid increase in weight and the deepening in color indicate a faster oxidation rate.
Based on the (2 2 0)’s full width at half maxima in Figure 4a and the (1 1 1) peak in Figure 4b,c, the grain size of the alloys were calculated using the Debye–Scherrer formula. The variation in grain size of the W220, W111, and WAS Ni–W alloys with temperature is displayed in Figure 6. The results show that the grain size of the as-deposited W220 is the largest (10.78 nm), and that of WAS is the smallest (9.08 nm). It can be concluded that the grain size of the Ni–W alloy decreases with the increase in W content. And, the grain size of the Ni–W alloy 2 h after annealing at 200 and 300 °C hardly changes. As the heat treatment temperature increases, the grains of the coatings grew to 20.60~39.82 nm at 600 °C. At 700 °C, the grain of WAS further coarsened to 30.82 nm, while W220 and W111 became small because of the formation of the new oxidation phase, presenting diffraction peaks at 2θ of 23.9 to 36.3°.

3.2. Microhardness

The microhardness of the three textured Ni–W alloy coatings is shown in Figure 7. The microhardness of the as-deposited W220, W111 and WAS samples ranged from 550 to 600 HV, being similar to the reported value of 460–740 HV [17]. It can be seen that the microhardness of WAS with the highest W content is highest within the entire annealed temperature range because of the solid solution strengthening effect of W [25]. Below 400 °C, the microhardness of the three textured Ni–W alloy coatings rapidly improved after heat treatment, being attributed to the elimination of residual stress and grain refinement crystallization [6,17].
At 400 °C, the microhardness of WAS reached 1028 HV. However, the microhardness for W220 was only 760 HV because its W content is lowest (see Table 2). Over 400 °C, the microhardness decreases along with the annealed temperature because of the coarsening crystal grains.

3.3. Corrosion Resistance

3.3.1. Tafel Curves

The corrosion resistance of the as-deposited and annealed Ni–W alloy coatings was characterized by Tafel in a 3.5 wt.% NaCl solution. For comparison, the matrix Q235 carbon steel disk was also used.
Figure 8 shows the Tafel curves of the three textured Ni–W alloy coatings of W220, W111 and WAS that were as-deposited and annealed at different temperatures in air for 2 h. And, the corrosion potential (Ecorr) and corrosion current density (Icorr) were obtained by the Tafel extrapolation method and are given in Table 4. It can be seen that the Ecorr saw a positive shift after these three textured Ni–W alloy coatings were electrodeposited onto the Q235 carbon steel disks, implying improved corrosion resistance [24]. The more the annealed temperature elevated, the more the Ecorr for all the samples is positively shifted. For the WAS sample, its Ecorr is significantly positively shifted after heat treatment at 600 °C, likely due to the completion of the structural transformation from amorphous to crystalline and the elimination of internal stress, which results in cracks in Ni–W obtained in an ammonium-citrate bath [17]. The ΔEcorr was calculated according to the difference between the samples and Fe blank electrode. The ΔEcorr is 165 mV, 139 mV and 112 mV for the as-deposited W220, W111 and WAS samples, respectively. That is, the value of ΔEcorr gradually decreases along with the increase in W content in Ni–W alloys. It was found that Icorr becomes high after coating Ni–W alloys, possibly because the galvanic coupling of Fe and Ni–W accelerate the corrosion of Fe in porous coatings. For every sample, the Icorr decreases along with the increase in annealed temperature. In general, the Icorr of the W220 sample is the lowest, which suggests that W220 possesses the best corrosion resistance. Alimadadi also reported that the corrosion resistance of single-phase Ni–W is superior to amorphous and dual-phase coated layers because of cracks or internal stress [26].
Improvement of corrosion resistance can be analyzed by the inhibition efficiency (IE, %), which was calculated according to the following equation (the results are shown in Table 4).
I E = I c o r r ,       F e I c o r r ,       N i W I c o r r ,       F e × 100 %
It can be seen that the IE decreased after the electrodeposition of the Ni–W alloy onto Q235 carbon steels, but IE can be increased by means of the heart treatment, and the biggest IEs are 85.8% for the W220 sample, 65.0% for the W111 sample and 55.2% for the WAS sample. As shown in Figure 5, the oxidation behavior occurs at temperatures of 400 °C or higher, which maybe decrease the porosity of the coating, and consequently, the results indicate that the W220 sample presents the best corrosion resistance.

3.3.2. EIS

The corrosion resistance of the Ni–W alloy coatings was also estimated by EIS spectra measured in the 3.5 wt. % NaCl solution, and the results are shown in Figure 9. All the Nyquist curves present a flat semicircle, suggesting that the Fe, as-deposited, and annealed Ni–W alloy coatings had undergone the same corrosion reaction process. However, the semicircular size of the as-deposited Ni–W alloy coatings is much larger than that of Fe, indicating the corrosion resistance is improved after electrodepositing Ni–W alloy coatings onto the Fe substrate. The semicircular arc became extraordinarily larger after heat treatment. Therefore, corrosion resistance can be significantly improved by annealing after electroplating.
Fitting through an equivalent circuit diagram shown in Figure 9d, the results are shown in Table 5. Here, RS is solution resistance, Rct is charge transfer resistance, and CPE1 is the double layer capacitance involved in the equivalent circuit. It can be seen that the Rct value increases from 1059 Ω cm2 for the Fe electrode to 10,743 Ω cm2 for the W220 sample, 5826 Ω cm2 for the W111 sample, and 2918 Ω cm2 for the WAS sample. The Rct value of the coatings was dramatically increased, along with annealed temperature. The maximum value of Rct is 43,267 Ω cm2 for the W220 sample, 28,178 Ω cm2 for the W111 sample, and 30,154 Ω cm2 for the WAS sample. The inhibition efficiencies were also calculated by Rct according to the following equation for IERct, and the results are shown in Table 5.
I E R c t = R c t ,       N i W R c t ,       F e R c t ,       N i W × 100 %
As expected, for all samples, the IERct values are all positive and increase with annealed temperature. The IERct of the as-deposited W220, W111 and WAS is 90.1%, 81.8% and 63.7%, respectively. Under the same condition, the WAS sample presents the biggest IERct, indicating the best corrosion resistance, which is consistent with that of the Tafel curves. For the three samples, the biggest value of IERct is 97.5% for W220, 96.2% for W111 and 96.5% for WAS.

3.4. HER

The electrocatalytic HER performance of the Ni–W alloy coatings was evaluated by LSV electrochemical technology in a 6 wt. % NaOH solution, and the results are shown in Figure 10. Compared with the Fe electrode, the W220, W111 and WAS alloys all showed improved electrocatalytic activity. With a driving current density of 100 mA cm−2, the HER potential required was −1.400 V for W220, −1.387 V for W111 and −1.413 V for WAS. As we know, the lower the potential is, the better the HER activity is. This result indicates that the HER activity of the (1 1 1) textured Ni–W alloy is superior to the (2 2 0) textured and amorphous samples. The LSV curves of the three samples annealed at 200, 400 and 600 °C are between that of Fe and the as-deposited Ni–W alloys; that is, the high HER overpotential is presented after heart treatment. Therefore, it is not beneficial to promote the HER property by means of heart treatment after electrodeposition.

4. Conclusions

In summary, Ni–W alloys with a (2 2 0) or (1 1 1) preferred orientation growth and amorphous structure were prepared by electrodeposition from a pyrophosphate bath. Low CW and Tb are conducive to the formation of a (220) plane preferential orientation, while on the contrary, a (111) orientation grew. The amorphous Ni–W alloys (WAS) with higher W content had the highest microhardness, being up to 1028 HV at 400 °C. The as-deposited and annealed Ni–W alloys with a (2 2 0) texture had better corrosion resistance, exhibiting a ΔEcorr of 329 mV, an Icorr of 0.81 uA cm−2 and an Rct of 43,267 Ω cm2. The initial HER overpotential of the (1 1 1) textured Ni–W alloy was the smallest—560 mV at 10 mA cm−2.

Author Contributions

Conceptualization and methodology: C.S., F.W. and J.D.; software: H.J. (Hui Ju), E.W. and C.Z.; validation., J.F., H.J. (Hongzong Jiang) and K.L.; formal analysis and investigation., Y.C., J.F., H.J. (Hui Ju) and K.L.; data curation: Y.L. and B.L.; writing—original draft preparation: J.D.; writing—review and editing: C.S. and F.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Provincial Natural Science Foundations of Hunan (2021JJ30184, 2021JJ30180, 2020JJ4243), the Scientific Research Fund of Hunan Provincial Education Department (20A108), and the Doctoral Research Fund of Hunan Institute of Engineering (09001003-21017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, C.; Chen, P.; Hu, B.; Xiang, Q.; Cen, Y.; Hu, B.; Liu, L.; Liu, Y.; Yu, D.; Chen, C. Porous nickel electrodes with controlled texture for the hydrogen evolution reaction and sodium borohydride electrooxidation. CrystEngComm 2020, 22, 4228–4237. [Google Scholar] [CrossRef]
  2. Lahiri, A.; Lakshya, A.K.; Guan, S.; Anguilano, L.; Chowdhury, A. Impact of different metallic forms of nickel on hydrogen evolution reaction. Scr. Mater. 2022, 218, 114829. [Google Scholar] [CrossRef]
  3. Eckold, P.; Niewa, R.; Hügel, W. Texture of electrodeposited tin layers and its influence on their corrosion behavior. Microelectron. Reliab. 2014, 54, 2578–2585. [Google Scholar] [CrossRef]
  4. Xu, Y.; Dai, Y.; Zhang, W.; Xia, T. Microstructure and texture evolution of electrodeposited coatings of nickel in the industrial electrolyte. Surf. Coat. Technol. 2017, 330, 170–177. [Google Scholar]
  5. Nasirpouri, F.; Sanaeian, M.R.; Samardak, A.S.; Sukovatitsina, E.V.; Ognev, A.V.; Chebotkevich, L.A.; Hosseinid, M.-G.; Abdolmaleki, M. An investigation on the effect of surface morphology and crystalline texture on corrosion behavior, structural and magnetic properties of electrodeposited nanocrystalline nickel films. Appl. Surf. Sci. 2014, 292, 795–805. [Google Scholar] [CrossRef]
  6. Allahyarzadeh, M.H.; Aliofkhazraei, M.; Rezvanian, A.R.; Torabinejad, V.; Sabour, R.A.R. Ni-W electrodeposited coatings: Characterization, properties and applications. Surf. Coat. Technol. 2016, 307, 978–1010. [Google Scholar] [CrossRef]
  7. Ramaprakash, M.; Deepika, Y.; Balamurugan, C.; Nagaganapathy, N.; Sekar, R.; Panda, S.K.; Rajasekaran, N. Pulse electrodeposition of nano-crystalline Ni-W alloy and the influence of tungsten composition on structure, microhardness and corrosion properties. J. Alloys Compd. 2021, 866, 158987. [Google Scholar] [CrossRef]
  8. Raveendran, M.; Neethu, A.; Chitharanjan, H. Development of Ni-W alloy coatings and their electrocatalytic activity for water splitting reaction. Physica B Condens. Matter. 2020, 597, 412359. [Google Scholar]
  9. Jameei, R.P.; Aliofkhazraei, M.; Barati, D.G. Ni-W nanostructure well-marked by Ni selective etching for enhanced hydrogen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 880–894. [Google Scholar] [CrossRef]
  10. Hong, S.H.; Ahn, S.H.; Choi, J.; Kim, J.Y.; Kim, H.Y.; Kim, H.-J.; Jang, J.H.; Kim, H.; Kim, S.-K. High-activity electrodeposited Ni-W catalysts for hydrogen evolution in alkaline water electrolysis. Appl. Surf. Sci. 2015, 349, 629–635. [Google Scholar] [CrossRef]
  11. Tang, J.; Niu, J.; Yang, C.; Rajendran, S.; Lei, Y.; Sawangphruk, M.; Zhang, X.; Qin, J. Twin boundaries boost the hydrogen evolution reaction on the solid solution of nickel and tungsten. Fuel 2022, 330, 125510. [Google Scholar] [CrossRef]
  12. Fraze, L. Shiny Science: A New Substitute for Hexavalent Chromium. Environ. Health Perspect. 2006, 114, A482–A485. [Google Scholar]
  13. Suvorov, D.; Gololobov, G.; Tarabrin, D.; Slivkin, E.; Karabanov, S.; Tolstoguzov, A. Electrochemical Deposition of Ni–W Crack-Free Coatings. Coatings 2018, 8, 233. [Google Scholar] [CrossRef]
  14. Su, C.; Sa, Z.; Liu, Y.; Zhao, L.; Wu, F.; Bai, W. Excellent Properties of Ni-15 wt.%W Alloy Electrodeposited from a Low-Temperature Pyrophosphate System. Coatings 2021, 11, 1262. [Google Scholar] [CrossRef]
  15. Popczyk, M.; Kubisztal, J.; Swinarew, A.S.; Waskiewicz, Z.; Stanula, A.; Knechtle, B. Corrosion Resistance of Heat-Treated Ni-W Alloy Coatings. Materials 2020, 13, 1172. [Google Scholar] [CrossRef]
  16. Hou, K.-H.; Chang, Y.-F.; Chang, S.-M.; Chang, C.-H. The heat treatment effect on the structure and mechanical properties of electrodeposited nano grain size Ni–W alloy coatings. Thin Solid Films 2010, 518, 7535–7540. [Google Scholar] [CrossRef]
  17. Su, C.S.; Zhao, L.F.; Bai, Y.; Tian, L.; Wen, B.; Guo, J.M. Microhardness Improvement of Ni-W/SiC Composite Coatings by High Frequency Induction Heat Treatment. J. Electrochem. Soc. 2019, 166, D301–D307. [Google Scholar] [CrossRef]
  18. Wang, H.; Sheu, H.; Ger, M.; Hou, K. The effect of heat treatment on the microstructure and mechanical properties of electrodeposited nanocrystalline Ni–W/diamond composite coatings. Surf. Coat. Technol. 2014, 259, 268–273. [Google Scholar] [CrossRef]
  19. Zhou, H.; Liao, Z.; Fang, C.; Li, H.; Feng, B.; Xu, S.; Cao, G.; Kuang, Y. Pulse electroplating of Ni-W-P coating and its anti-corrosion performance. Trans. Nonferrous Met. Soc. China 2018, 28, 88–95. [Google Scholar] [CrossRef]
  20. He, F.J.; Fang, Y.Z.; Jin, S.J. The study of corrosion–wear mechanism of Ni–W–P alloy. Wear 2014, 311, 14–20. [Google Scholar] [CrossRef]
  21. Li, B.; Zhang, W. Microstructural, surface and electrochemical properties of pulse electrodeposited Ni–W/Si3N4 nanocomposite coating. Ceram. Int. 2018, 44, 19907–19918. [Google Scholar] [CrossRef]
  22. Luo, H.; Leitch, M.; Zeng, H.; Luo, J.-L. Characterization of microstructure and properties of electroless duplex Ni-W-P/Ni-P nano-ZrO2 composite coating. Mater. Today Phys. 2018, 4, 36–42. [Google Scholar] [CrossRef]
  23. Chou, C.-C.; Huang, P.-C.; Dai, H.-B. An Analysis of the Performance Characteristics of Different Cylinder Temperatures for Ni-W and Ni-W-BN(h) Plated Piston Rings in an Air-Cooled Engine. Energies 2022, 15, 1026. [Google Scholar] [CrossRef]
  24. Sinem, E.; Ürgen, M. Oxidation behavior of electroless Ni–P, Ni–B and Ni–W–B coatings deposited on steel substrates. Surf. Coat. Technol. 2015, 265, 46–52. [Google Scholar]
  25. Yin, D.; Marvel, C.J.; Cui, F.Y.; Vinci, R.P.; Harmer, M.P. Microstructure and fracture toughness of electrodeposited Ni-21 at. % W alloy thick films. Acta Mater. 2018, 143, 272–280. [Google Scholar] [CrossRef]
  26. Alimadadi, H.; Ahmadi, M.; Aliofkhazraei, M.; Younesi, S.R. Corrosion properties of electrodeposited nanocrystalline and amorphous patterned Ni-W alloy. Mater. Des. 2009, 30, 1356–1361. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of Ni–W coatings prepared at current densities of 4, 6, 8 and 10 A dm−2 (a) at CW = 30 g L−1, Tb = 40 °C, (b) at CW = 60 g L−1, Tb = 40 °C, (c) at CW = 30 g L−1, Tb = 60 °C, and (d) at CW = 60 g L−1, Tb = 60 °C.
Figure 1. The XRD patterns of Ni–W coatings prepared at current densities of 4, 6, 8 and 10 A dm−2 (a) at CW = 30 g L−1, Tb = 40 °C, (b) at CW = 60 g L−1, Tb = 40 °C, (c) at CW = 30 g L−1, Tb = 60 °C, and (d) at CW = 60 g L−1, Tb = 60 °C.
Coatings 13 01519 g001
Figure 2. The relationship between current density and W content in Ni–W coatings prepared at different levels of Tb from bathes containing (a) CW = 30 g L−1 and (b) CW = 60 g L−1.
Figure 2. The relationship between current density and W content in Ni–W coatings prepared at different levels of Tb from bathes containing (a) CW = 30 g L−1 and (b) CW = 60 g L−1.
Coatings 13 01519 g002
Figure 3. Ni–W alloy SEM images of (a) W220, (b) W111 and (c) WAS samples with (2 2 0) texture, (1 1 1) texture and amorphous structure, respectively, and (d) the thickness of W220, W111 and WAS samples.
Figure 3. Ni–W alloy SEM images of (a) W220, (b) W111 and (c) WAS samples with (2 2 0) texture, (1 1 1) texture and amorphous structure, respectively, and (d) the thickness of W220, W111 and WAS samples.
Coatings 13 01519 g003
Figure 4. The XRD pattern evolution of (a) W220, (b) W111 and (c) WAS along with the annealed temperature.
Figure 4. The XRD pattern evolution of (a) W220, (b) W111 and (c) WAS along with the annealed temperature.
Coatings 13 01519 g004
Figure 5. Differential scanning calorimetry (DSC) and thermogravimetry (TG) curves of (a) W220, (b) W111 and (c) WAS samples. The inserts are the photographs of carbon steel disks coated with Ni–W alloys and annealed at 200, 300, 400, 500, 600 and 700 °C.
Figure 5. Differential scanning calorimetry (DSC) and thermogravimetry (TG) curves of (a) W220, (b) W111 and (c) WAS samples. The inserts are the photographs of carbon steel disks coated with Ni–W alloys and annealed at 200, 300, 400, 500, 600 and 700 °C.
Coatings 13 01519 g005
Figure 6. The effect of annealed temperature on the crystallite size of three textured Ni–W alloys.
Figure 6. The effect of annealed temperature on the crystallite size of three textured Ni–W alloys.
Coatings 13 01519 g006
Figure 7. The microhardness of the three textured Ni–W alloy coatings annealed at different temperatures from 200 to 700 °C in air for 2 h.
Figure 7. The microhardness of the three textured Ni–W alloy coatings annealed at different temperatures from 200 to 700 °C in air for 2 h.
Coatings 13 01519 g007
Figure 8. Tafel curves of three textured Ni–W alloy coatings of (a) W220, (b) W111 and (c) WAS as-deposited and annealed at different temperatures in air for 2 h. For comparison, aQ235 carbon steel disk is also given.
Figure 8. Tafel curves of three textured Ni–W alloy coatings of (a) W220, (b) W111 and (c) WAS as-deposited and annealed at different temperatures in air for 2 h. For comparison, aQ235 carbon steel disk is also given.
Coatings 13 01519 g008
Figure 9. Nyquist plots of as-deposited and annealed (a) W220, (b) W111 and (c) WAS; (d) is an equivalent circuit.
Figure 9. Nyquist plots of as-deposited and annealed (a) W220, (b) W111 and (c) WAS; (d) is an equivalent circuit.
Coatings 13 01519 g009
Figure 10. LSV curves of (a) W220, (b) W111 and (c) WAS in 6 wt. % NaOH solution.
Figure 10. LSV curves of (a) W220, (b) W111 and (c) WAS in 6 wt. % NaOH solution.
Coatings 13 01519 g010aCoatings 13 01519 g010b
Table 1. The chemical components of the Q235 steel.
Table 1. The chemical components of the Q235 steel.
ComponentC/%Si/%Mn/%P/%S/%Fe/%
Q235 steel0.12–0.2<0.30.3–0.7<0.045<0.045Remaining
Table 2. Bath compositions and plating conditions for textured Ni–W alloys.
Table 2. Bath compositions and plating conditions for textured Ni–W alloys.
Chemicals/ParametersValues (g L−1)
NiSO4·6H2O12
NaWO4·2H2O30, 60
Na4P2O7·10H2O40
H3PO4 (85%)10
NH3·H2O (25%–28%)~20
Bath temperature40 °C, 60 °C
Current density4–10 A dm−2
pH8.5–9.0
Table 3. Electrodeposition parameters of three textured Ni–W alloys.
Table 3. Electrodeposition parameters of three textured Ni–W alloys.
NameW wt. %OrientationCWTb/°CCurrent/A dm−2
W22013.85(2 2 0)30408
W11129.80(1 1 1)6040
WAS36.75Amorphous6060
Table 4. Corrosion resistance of three textured Ni–W alloy coatings.
Table 4. Corrosion resistance of three textured Ni–W alloy coatings.
SampleT/°CIcorr/uAcm−2Ecorr/mV vs. SHEΔEcorr/mVIE/%
Fe-5.71−3700-
W220As-deposited13.2−205165−131.2
2005.77−174196−1.1
4002.55−6430655.3
6000.81−4132985.8
W111As-deposited16.27−231139−184.9
20010.93−207163−91.4
4002.00−12724365.0
6003.22−6830243.6
WASAs-deposited16.82−258112−194.6
20011.49−236134−101.2
40011.41−223147−99.8
6002.56−9627455.2
Table 5. EIS Fitting results of three textured Ni–W alloys according to the equivalent circuit shown in Figure 9d.
Table 5. EIS Fitting results of three textured Ni–W alloys according to the equivalent circuit shown in Figure 9d.
SampleT/°CRs/Ω cm2CPE-T/uF cm−2CPE-PRct/Ω cm2IERct/%
Fe-5.9812640.7531059-
W220As-deposited6.564190.6910,74390.1
2005.262230.7412,06591.2
4006.121100.7818,01294.1
6005.85780.8943,26797.5
W111As-deposited6.2841060.72582681.8
2006.7481370.77651183.7
4005.07490.88632583.2
6006.043460.8828,17896.2
WASAs-deposited5.635300.78291863.7
2004.5252190.77475177.7
4005.8211370.78808386.9
6005.9140.66330,15496.5
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

Deng, J.; Li, K.; Fu, J.; Li, B.; Jiang, H.; Ju, H.; Wang, E.; Zhang, C.; Liu, Y.; Chen, Y.; et al. Preparation and Properties of Textured Ni–W Coatings Electrodeposited on the Steel Surface from a Pyrophosphate Bath. Coatings 2023, 13, 1519. https://doi.org/10.3390/coatings13091519

AMA Style

Deng J, Li K, Fu J, Li B, Jiang H, Ju H, Wang E, Zhang C, Liu Y, Chen Y, et al. Preparation and Properties of Textured Ni–W Coatings Electrodeposited on the Steel Surface from a Pyrophosphate Bath. Coatings. 2023; 13(9):1519. https://doi.org/10.3390/coatings13091519

Chicago/Turabian Style

Deng, Jiyu, Kunpeng Li, Jianglong Fu, Bing Li, Hongzong Jiang, Hui Ju, Erli Wang, Changke Zhang, Yangyang Liu, Yan Chen, and et al. 2023. "Preparation and Properties of Textured Ni–W Coatings Electrodeposited on the Steel Surface from a Pyrophosphate Bath" Coatings 13, no. 9: 1519. https://doi.org/10.3390/coatings13091519

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