Next Article in Journal
Microwave Absorption Properties of Fe3O4 Particles Coated with Al via Rotating Magnetic Field Method
Next Article in Special Issue
Plasma Electrolytic Oxidation Ceramic Coatings on Zirconium (Zr) and ZrAlloys: Part I—Growth Mechanisms, Microstructure, and Chemical Composition
Previous Article in Journal
Physical Vapour Deposited Biomedical Coatings
Previous Article in Special Issue
Aluminide Thermal Barrier Coating for High Temperature Performance of MAR 247 Nickel Based Superalloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plasma Electrolytic Oxidation Ceramic Coatings on Zirconium (Zr) and Zr-Alloys: Part-II: Properties and Applications

1
Center for Advanced Materials Research, University of Texas at El Paso, 500 W. Univ. Ave., El Paso, TX 79968, USA
2
Environmental Science and Engineering, University of Texas at El Paso, 500 W. Univ. Ave., El Paso, TX 79968, USA
3
Department of Mechanical Engineering, University of Texas at El Paso, 500 W. Univ. Ave., El Paso, TX 79968, USA
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(6), 620; https://doi.org/10.3390/coatings11060620
Submission received: 13 May 2021 / Accepted: 18 May 2021 / Published: 22 May 2021

Abstract

:
A plasma electrolytic oxidation (PEO) is an electrochemical and eco-friendly process where the surface features of the metal substrate are changed remarkably by electrochemical reactions accompanied by plasma micro-discharges. A stiff, adhesive, and conformal oxide layer on the Zr and Zr-alloy substrates can be formed by applying the PEO process. The review describes recent progress on various applications and functionality of PEO coatings in light of increasing industrial, medical, and optoelectronic demands for the production of advanced coatings. Besides, it explains how the PEO coating can address concerns about employing protective and long-lasting coatings with a remarkable biocompatibility and a broad excitation and absorption range of photoluminescence. A general overview of the process parameters of coatings is provided, accompanied by some information related to the biological conditions, under which, coatings are expected to function. The focus is to explain how the biocompatibility of coatings can be improved by tailoring the coating process. After that, corrosion and wear performance of PEO coatings are described in light of recognizing parameters that lead to the formation of coatings with outstanding performance in extreme loading conditions and corrosive environments. Finally, a future outlook and suggested research areas are outlined. The emerging applications derived from paramount features of the coating are considered in light of practical properties of coatings in areas including biocompatibility and bioactivity, corrosion and wear protection, and photoluminescence of coatings

1. Introduction

Plasma electrolytic oxidation (PEO) is a promising electrochemical technique to form ceramic coatings using plasma-assisted oxidation on a broad range of metals, such as Mg, Ti, and Zr [1,2,3]. The significant merit of this technique is to fabricate protective inorganic layers grown from the substrate with the participation of species from the electrolyte. Forming PEO layers on valve metals boosts surface performances in many applications by enhancing the structural reliability and extending the longevity of functional properties. During the PEO process, the plasma state following extreme voltage and temperature transmits to the surface of the anodic substrate in an alkaline electrolyte and produces inorganic layers with outstanding adhesive strength to the underlayer [4,5]. As asserted unanimously, the PEO technique forms an environmentally friendly, wet-based coating implementing alkaline electrolytes as the primary media and providing conditions for the incorporation of inorganic additives, including silicate, phosphate, aluminate, or organic additives, including sodium oxalate, glycerol to facilitate the physicochemical attraction during the process via electrochemical reactions [6].
We explained the paramount roles of the electrical, compositional variables in determining the characteristics of coatings formed during the PEO process. These variables, such as current mode and density, waveform, and duty cycle, would considerably influence the alteration of coatings’ surface and composition features. Therefore, earlier studies have been reported to shed light on the effects of PEO parameters on microdischarge characteristics and obtained features of inorganic layers on the Zr and Zr-alloys. Tailoring process variables during the PEO results in improving the applicability and enhancing the functionality of coatings in various aspects. For instance, the excellent corrosion properties of Zr-alloys were attained via controlling both electrical variables and electrolyte composition, which could, in turn, influence the spatial, lifetime, and electrical characteristics of the PEO coatings [7,8,9,10,11,12,13]. To the best of our knowledge, no comprehensive review has discussed the applications of the PEO method to prevent degradation, corrosion, and wear, and to improve the biocompatibility of Zr and Zr-alloy extensively. The recent advances in biological responses of Zr implants using the PEO coatings have been reviewed by Fattah-alhosseini’s group recently [3]. Furthermore, we presented a detailed review of PEO coatings in Part-I of this issue, where the attention given was on deriving a deeper fundamental understanding of the PEO growth mechanisms and the effect of process parameters on transient discharge behavior at breakdown, initiation, and growth of the oxide layer, and on the incorporation of species from electrolyte. Part-I highlights the fundamental microstructural aspects associated with structural defects, phase transformation, and the role of additives. However, a detailed report on the properties and application of PEO coatings is still missing.
Considering the scarcity of reviews on the PEO coating of Zr and Zr-alloys as a specific group, a comprehensive review is strongly required to provide the integrated perspective on the applicability and performance of PEO coatings respective to current demands in industries, medication, and science. Accordingly, the present review provides the functional features of PEO to approach studying the applicability of coatings with special emphasis on imperative phenomena and applications of interest, such as biocompatibility in medical implantation, extensive endurance of coatings under applied loadings, outstanding protection of underlayer in exposure to corrosive environments, and a broad photoluminescence range. We describe the general picture of structural reliabilities in correspondence with tribological and electrochemical reactions, and the emerging frame of functional properties in a broad range of emitted light is also provided. Finally, future outlook and prospective progress in synergy with enhancing both pre- and post-treatments are given to provide advanced functionality of PEO coatings are outlined.

2. Biocompatibility of PEO Coatings

There is a significant desire to substitute Ti with Zr for surgical implant materials because of lower Young’s modulus (92 GPa) compared to titanium (110 GPa), and hence, lower risk for implantation failure due to stress shielding [14,15]. The biocompatibility of Zr is related to the formation of an inherent oxide layer of zirconia, which is bio-inert and restricts its integration with bone tissue during implantation [16]. Therefore, direct bonding with bone and stimulating new bone formation on the Zr surface is not feasible. Besides, the inherent oxide layer with 2 to 5 nm thickness is prone to rupture, erosion, and wear once Zr is employed for high load-bearing implant applications [17,18]. To address these drawbacks, several surface modifications are implemented to boost the bioactivity and corrosion resistance of Zr without compromising its biocompatibility [18,19,20,21,22]. Out of numerous techniques, the PEO process has demonstrated a promising technique to develop firmly adherent, porous, crystalline, relatively rough, and thick oxide coatings on Zr in environmentally friendly alkaline based electrolytes.
Generally, an implant needs to demonstrate outstanding surface features, such as superior bioactivity and biocompatibility. It is expected that implants applicable to dental and orthopedic surfaces demonstrate the capability to facilitate new bone formation and bond steadily and firmly with bone following implantation [23]. The procedure to examine the apatite-forming ability of PEO coatings and Zr substrate is that they are immersed in simulated body fluid (SBF) medium for 14 days, as shown in Figure 1a. After that, the surface can be characterized using techniques, such as scanning electron microscopy-energy dispersive spectroscopy (SEM-EDS), for studying mineralization of the formed apatite layer on the coating surface and checking the elemental composition of the layer. The SEM image illustrated in Figure 1a showed formation of apatite compounds after immersion in SBF for 14 days. Precipitation of compounds composed of Ca and P over the surface was considered the indication of apatite formation. It was evident that sphere-shaped deposits covered the entire surface of all PEO coatings. However, the precipitated layer on the bare Zr surface was not uniform. Elemental analysis showed that the Ca/P ratios of precipitates on PEO coatings were in the range of 1.53 to 1.65, which is close to the stochiometric Ca/P ratio of 1.67 for hydroxyapatite [24].
Yan et al., took advantage of a post-treatment using ultraviolet (UV) irradiation to change the surface chemistry of the PEO coating. It turned out that strong UV irradiation caused the formation of more basic Zr-OH groups on the surface, and, therefore, the hydrophilic nature of the surface improved significantly [25]. Han et al., produced PEO coating containing Ca and P and studied the in-vitro bioactivity, and dominant osteoblast response of the film as a function of the applied voltage ranging from 400 to 500 V [26]. The PEO coating with a predominantly t-ZrO2 phase could facilitate the accumulation of a more significant quantity of CaO and P compounds at higher applied voltage, while the apatite-forming ability of coatings was also boosted. A facile method to examine bioactivity and biocompatibility of PEO coatings was to study the proliferation and growth feasibility of the attached osteoblast cells on the surface in an alkaline phosphate environment [26]. It was deduced that doping ZrO2 with Ca2+ ions could improve the accumulation of hydroxyapatite (HA) on the ZrO2 surface. Yan et al., also worked on enhancing the apatite-forming ability by activating the PEO surface using acidic or alkaline solutions. Significant differences in formation of HA on the surface were noticed between the etched PEO surface and intact surface, where HA formed after a short time on the treated surface [27]. Zhang et al., also introduced HA nanorods into the PEO coating via a process in which HA nanorods were prepared via hydrothermal treatment over the PEO coating [28]. The produced calcium partially stabilized zirconia (Ca-PSZ) coating became covered with HA nanorods and showed greater hydrophilicity and apatite-forming ability. That coating was usually prepared in multistep methods or did not contain a crystalline HA layer on the substrates [27,28]. Cengiz et al., prepared a crystalline HA layer using a single-step PEO process on zirconium with enhanced bioactivity [29]. Composition characterization of PEO coatings showed that hydroxyapatite (HA) formed directly in the coating along with calcium zirconium oxide (Ca0.134Zr0.86O1.86) Ca-PSZ and m-ZrO2. The formation mechanism of the PEO coating with three layers is displayed in Figure 1b. It was found that Ca and P significantly accumulated in the outer region of the coating, while the inner region was rich in Zr and O. The reaction between pre-formed ZrO2 and Ca2+ ions from the electrolyte resulted in formation of the Ca–Zr–O phase in the transition region. After that, high energy sparking regimes during the PEO process could induce HA layer formation, which could be initiated by complex reactions between Ca2+ and PO 4 3 ions on the anode surface. Thus, coatings with higher wettability could form during the PEO process. Figure 2a,b compares the wettability between the Zr substrate and the modified PEO coating. The formation of hydrophilic HA coating enhanced the wettability and reduced contact angle values in comparison with a bare Zr surface [29].
The crystallinity and morphology of HA phases could play a vital role in cell growth and cell activity [30,31]. Non-crystalline calcium phosphates were not favorable as they might be dissolved in human body fluids, implying instability for implantation [30,32]. Fu et al., studied 3T3 cell proliferation as a marker to test the biocompatibility of PEO coatings. They were able to measure 3T3 proliferation by MTS assay for PEO coatings formed in electrolytes with different KOH concentrations. The cell proliferation ability was enhanced significantly after seeding 3T3 cell on the PEO surface, however, the growth rate of 3T3 cells were similar regardless of their formation electrolytes [33].
In contrast to the many efforts to increase biocompatibility of PEO coatings on Zr, one of the other aims for surface improvements was to decrease bacterial activity or colonization in the vicinity of implants. Antibacterial activity of PEO coatings on Ti was enhanced by adding silver (Ag), thereby incorporating as an antibacterial agent [34,35]. Fidan et al., produced PEO coatings with silver acetate additives to halt bacterial activities of methicillin-resistant Staphylococcus aureus (MRSA) on the coatings [36]. Nanoparticles of Ag at concentration below toxicity levels could penetrate environments containing bacteria and expedite the release of Ag+ ions, due to the greater specific surface. As shown in Figure 2c, Ag+ ions demonstrated outstanding antibacterial activity against MRSA as one of the dominant reasons for infection after implantation. Durdu and his co-worker also worked on antibacterial properties of Ag-based bioceramic coating, where they deposited a thin layer of Ag hydrothermally on PEO coated Zr. They found that Ag distributed uniformly on the surface and that the final surface showed hydrophobic nature, in which the bacterial adhesion of the final surface reduced significantly compared to the condition without the Ag thin layer [37].
Aktuğ et al., produced PEO coatings composed of c-ZrO2, perovskite-CaZrO3, and hydroxyapatite (HA) in a single step for 5, 10, and 15 min [38]. They aimed to study the effect of PEO treatment time on the fabrication of a bioceramic-based hydroxyapatite coating. They studied the adhesion of Gram-positive and Gram-negative bacteria, including E. coli, P. aureginosa, P. putida, B. subtilis, S. aureus, and E. faecalis on the PEO coating. Hemolysis and MTT assays were used to examine the biocompatibility of coatings. Besides, the bioactivity was also tested using an in vitro immersion test for up to 28 days. Authors found that PEO surfaces demonstrated particular bioselectivity and modified cell–surface interactions. The microbial adhesion decreased with treatment time due to alteration of surface properties in thicker coatings. The difference in cell adhesion activity for Gram-negative and Gram-positive bacteria could be related to difference in cell surface hydrophilicity obtained by the higher amount of HA in prolonged processes [38]. Later, the same research group continued working on the production of PEO coatings containing HA composition to accelerate their interaction and bioactivity once they would be implanted [39].
Cengiz et al., studied the relation between microdischarge types and bioactivity of the PEO coatings for different treatment times ranging from 2.5 to 30 min [40]. The surface characterization revealed that the tetragonal phase formed as the dominant phase during all process times in electrolyte-containing Ca(CH3COO)2·H2O and C3H7Na2O6P·5H2O. It was found that the amorphous HA formed during the PEO process on the coating surface facilitated the formation of HA crystals once coatings were immersed in SBF solution [40]. The combination of the PEO process and electrophoretic deposition (EPD) was also implemented to form bioactive ZrO2/HA composite film [41]. Negatively charged HA nanoparticles were added to phosphate electrolyte to participate in a single-step PEO-EPD process for 2 to 6 min. The coating characterization showed that the uniform and dense ZrO2/HA coating consisted of c-ZrO2 and m-ZrO2 nanocrystalline phases. HA nanoparticles were attracted to the discharge channels due to their charge and entrapped in channels because of the EPD process, while coatings supported greater wettability. After immersion into SBF for 8 days, an apatite layer covered throughout the surface of ZrO2/HA coating, indicating considerable modification of in-vitro bioactivity. Besides, it was revealed that the human osteosarcoma could attach, adhere, and propagate considerably on the surface of the ZrO2–HA layer, implying the possibility of its implementation as orthopedic implant material [41].
Durdu et al., worked on thermal deposition of the Zn thin film on the PEO coating produced in electrolyte containing calcium acetate and β-CaGP salt [42]. The characterization of phase composition of coatings revealed that c-ZrO2, CaZrO3, and HA formed during the PEO process. The deposition of a uniform Zn layer on the entire surface caused super hydrophilicity compared to the undeposited surface. After immersion in SBF for 10 days, a compact and uniform bone-like apatite formed on the Zn-bioceramic surface. Microbial adherence tests for Gram-negative and Gram-positive bacteria exhibited that the antibacterial activity of the Zn-deposited surface was significantly strong against both types of bacteria, indicating enhanced bioactivity and antibacterial properties for the Zn-deposited PEO coating [42]. Zhang et al., fabricated HA nanorod-patterned zirconia coatings via a hybrid approach [43]. They found that the HA nanorod-patterned coating increased the feasibility of protein absorption and significantly modified the adhesive and proliferative properties of fibroblasts in comparison to untreated coating, suggesting a potential application for percutaneous implants to facilitate the connection to the skin [43]. Recently, Cu was also deposited onto the PEO surfaces without inducing any alteration in surface morphology by using thermal evaporation [44]. Surface characterization revealed the formation of meta-stable calcium zirconium oxide, cubic zirconia, and perovskite calcium zirconate phase. After completing the coating process, coatings with and without Cu top-layer were immersed in simulated body fluid to evaluate their in-vitro bioactivity and antimicrobial properties for Gram-positive and Gram-negative bacteria. Deposition of Cu hydrothermally on the PEO surface enhanced hydrophobic features, where both types of bacteria had a shorter microbial adhesion time to the surface of Cu-deposited PEO coating, thereby suggesting the biomedical application of the surface-modified PEO coating [44].
Very recently, the effect of pore size was studied in PEO coatings with respect to bioactivity and cell interaction properties [45]. Cengiz et al., prepared PEO coatings with various surface morphologies in terms of pore size and surface roughness to investigate in vitro bioactivity and cell interaction features after prolonged immersion in SBF. The in vitro bioactivity evaluations revealed the pore size of the coating significantly affected the growth rate of the HA layer on the coating surface. Remarkably, increasing the pore size promoted HA layer formation in SBF solution, in which higher roughness created by the PEO process increased the chance of HA nucleation, due to boosting the ionic activity of the coating surface [45].

3. Corrosion Resistance Properties

Nuclear grade Zr-alloys like Zircaloy-4 alloy are subjected to neutron irradiation as well as severe conditions of high temperature and high-pressure flowing water [46]. The harsh, aggressive environment exacerbates the corrosion and hydrogen absorption, thus causing fast deterioration of the mechanical strength of structural components [47,48]. Therefore, it is necessary for the nuclear industry to seek methods of curbing the corrosion rate of Zr alloys. On the other hand, Zr and its alloys have recently found a promising application as medical implants. The tendency of implant materials to be subjected to electrochemical reactions within body fluids is crucial to understanding their stability in the human body. Several techniques can be implemented to mimic the situation in the body and study the behavior of coated or noncoated Zr surfaces. Open circuit potential (OCP), Tefel extrapolation, potentiodynamic polarization (PDP), and electrochemical impedance spectroscopy (EIS) are the typical tests carried out on the PEO coated Zr samples to examine their corrosion performance in simulated body fluid (SBF) or other corrosive environments.
The PEO process is identified as a relatively fast, facile, and uncomplicated technique to produce functional ceramic coatings for corrosion-resistant applications on valve metals such as Ti and Zr. It is evident that the corrosion protection of the bare Zr substrate is associated with forming a native oxide layer on its surface. However, this oxide layer is not thick enough (i.e., in the order of a few tens of nanometer) and can be disappeared or degraded in the early stage of load bearing in prolonged applications like medical implants. This ends up raising the corrosion rate, thus reducing the efficacy and service time of the Zr implant. Although PEO coating is an effective method for lowering the corrosion rate due to blocking the charge-transfer mobility across the substrate–environment interface, the PEO coating suffers from its inherited porous nature, where deep pores provide an open path for penetration of aggressive ions and reduction in the overall protection efficiency of coating [9,10,11,12,13,24,49,50,51,52]. The PEO coating can form in various electrolytes, such as silicate, aluminate, phosphate, and so on, but all suffer from pancake-like surface features, pores, and cracks [53,54,55]. The rough and porous surface morphology did not disappear even for procedures without creating pancake-like structures [13,56]. Xue et al., explored the corrosion performance of PEO coating formed in silicate electrolyte using potentiodynamic polarization (PDP) tests [57]. They found that corrosion potentials shifted to more noble potentials and that the corrosion current density dropped several orders of magnitude due to the PEO coating. They claimed thickening the PEO coating impacted the corrosion rate, due to the higher number of cracks [57].
Sandhyarani et al., conducted a series of corrosion experiments to examine the corrosion performance of PEO coatings produced in the phosphate electrolyte with different treatment times. They found that the recorded OCPs of PEO coatings shifted to more noble potentials with a remarkable decrease in icorr compared to that of the untreated Zr sample [12]. No pitting corrosion was observed for the PEO coating produced below 6 min treatment time. However, the PEO coating produced with the longer treatment time (8 min) suffered significant pitting corrosion, which was attributed to numerous cracks on the coated surface. The optimum corrosion performance in SBF environment was recorded for the Zr sample coated up to 6 min, while precipitates were mainly composed of Ca–P rich compounds [12].
The corrosion performance of the PEO coating produced on Zircaloy-4 in the silicate electrolyte using AC regime was tested in nitric acid solution [11]. The authors studied the influence of the frequency of the PEO process on the corrosion resistance of the formed coatings. Electrochemical studies revealed that the performance of coatings produced at both 50 and 100 Hz for 5 and 30 min were similar. No significant difference in the corrosion resistance was found for coatings produced at different frequencies, thus confirming the effect of similar coating morphologies. The Tafel test showed the current density declined remarkably for the PEO coating compared to the bare zirconium, and no noticeable difference was found between coatings with different thicknesses. The authors, therefore, deduced that the main protection was supported by the barrier layer at the base of the coatings [11].
The corrosion performance of PEO coatings produced in Na2SiO3·9H2O and KOH solution with 1, 3, 10, and 30 min treatment times was tested in 0.5 M NaCl solution. Among PEO coatings, the coating with 10 min treatment time demonstrated the best corrosion protection properties. Increased porosity was the determining factor in decreasing the corrosion resistance of the PEO coating with 30 min treatment time [51].
Sandhyarani et al., also evaluated the effect of electrolyte composition on the corrosion performance of PEO coatings. They recorded OCP of PEO coatings formed in phosphate and aluminate electrolytes for 4 h and showed OCP values of all PEO coatings tended to more noble potentials, compared to Zr substrate, indicating higher thermodynamic stability of oxide films. At the beginning of immersion, OCP values for PEO coatings showed less noble potential and declined slightly. This instability at the early time of immersion could be associated with an increased activity because of penetration of corrosive solution from surface defects and pores to the metal–oxide interface. The oxide film formed in phosphate and KOH electrolyte showed the highest noble potential values over the entire period of immersion time, which can be related to the modified coating morphology of being pore-free and much smoother than other coatings. However, the film formed in silicate and KOH electrolyte showed higher apatite-forming ability because of superior wettability.
The evaluation of corrosion performance of the PEO composite coatings reiterated the importance of adding nanoparticles, in which the PEO composite coating produced with Ce2O3 addition could reduce the corrosion current density about 103 times compared to the simple PEO coating [9]. Figure 3a shows the polarization curves for the Zr substrate, the simple PEO, and the PEO composite coating with Al2O3, CeO2, or ZrO2 nanoparticles, termed as PA, PC, and PZ, respectively. Besides, the composite coating of ZrO2–SiC produced in phosphate electrolyte exhibited superior corrosion resistance compared to coatings produced in aluminate and silicate electrolytes, due to the modified surface morphology of the coating [58].
Wang et al., conducted a comparative study on two types of PEO coatings produced on Zircaloy-4 alloy in silicate and phosphate electrolyte. The authors worked on the PDP behaviors and electrochemical impedance spectroscopy (EIS) measurements of PEO coatings immersed in LiOH solution for different time intervals to study the prolonged corrosion performance [59]. The surface morphology of the two coatings showed distinct differences, where the PEO coating formed in silicate electrolyte was more compact, and the population of pores and cracks was fewer than the PEO coating formed in phosphate electrolyte. The PEO coating formed in phosphate electrolyte suffered from localized peeling regions caused by stress mismatching of the considerable volume change owing to the transition from tetragonal to monoclinic phases. Figure 3b illustrates the PDP analysis of PEO coatings and the bare Zr substrate, which displayed no passive region for the PEO coatings as a remarkable difference with the bare sample. The corrosion current density is three and four orders of magnitude lower than that of the bare electrode for PEO coatings formed in phosphate and silicate electrolyte, respectively.
EIS measurements were performed for both coatings immersed in LiOH solution, and various equivalent electrical circuits were suggested to simulate results. Figure 4a,b shows the Nyquist plots of EIS results for silicate and phosphate coatings, and Figure 4c,d shows the suggested electrical circuits for prolonged immersions. It was evident that the silicate coating with modified surface morphology could endure the aggressive condition greater and protect zirconium more significantly in the long-term immersion process.
Li et al., mixed graphene oxide (GO) particles into silicate and a mixture of phosphate and silicate electrolytes to produce the PEO composite coatings of zirconia with graphene on N36 Zr alloy [8]. They studied corrosion and fretting behavior of the composite coating and investigated the influence of GO addition on improving the corrosion resistance. The authors reported the positive role of GO particles in enhancing the anti-corrosion behavior of PEO coating, in which the coating formed with 0.1 g/L GO at silicate electrolyte exhibited the best corrosion performance. The GO particles developed a good barrier against penetration of aggressive ions and blocked access to the substrate [8]. Very recently, the same authors compared the corrosion performance of PEO composite coating with each of Al2O3, MoS2, CeO2, and graphene oxide (GO) on Zr alloy [7]. The positive effect of adding nanoparticles was to lower the corrosion current density with more noble corrosion potential due to impeding aggressive ions before they reach the substrate.
Wu et al., combined the PEO process with pulsed laser deposition (PLD) to enhance the corrosion protection of composite film formed on Zr-4 alloy [60]. PLD was used to form a Cr/CrN/Cr2O3 film covering the PEO coating to boost the corrosion resistance of the film on Zr alloy. The EIS study confirmed improving the corrosion resistance via enlarging the polarization resistance, suggesting contribution of dense Cr/CrN/Cr2O3 film on the corrosion performance of ZrO2 film though blocking ion attacks more effectively.

4. Wear and Tribology of the PEO Coatings

Zr alloys demonstrated low neutron absorption coefficients and high yield strength at higher temperatures, which led to their selection for nuclear applications, such as fuel tubes and coolant channel materials in water-cooled power reactors [10,46,61,62]. Furthermore, Zr alloys suffer from low wear resistance in comparison with other nuclear materials, such as stainless steels and nickel alloys [63], while the fretting wear on tubes is an important concern in pressurized water reactors [63,64]. Besides, Zr alloys also demonstrated potential for biomedical applications, and thus, improving wear resistance can allow Zr-based implants to replace Ti-based implants. The dry sliding wear performance of the PEO coating can be related to main properties, including microstructure, composition, thickness of coatings, and the sliding procedures, which are related to the applied load and sliding time. Figure 5a shows the wear scar of the PEO coating on Zircaloy-2 in silicate electrolyte for 20 min. Due to the sliding wear test, it was evident that the coating was completely peeled off after 600 s, and the ploughing effects were visible on the substrate. As shown in Figure 5b,c, the PEO coating formed in aluminate electrolyte for 10 min is not worn out fully even after 30 min of sliding wear test [53]. The weak performance of the PEO coating formed in silicate electrolyte could be attributed to numerous cracks and pores in the coating, while the PEO coating formed in aluminate electrolyte showed greater wear performance due to the coating’s compactness and higher t-ZrO2 content.
Cheng et al., performed dry sliding wear tests using an applied load of 10 N for coatings produced for 5, 10, and 30 min in dilute electrolyte of aluminate and KOH [54]. The friction coefficient of the PEO coating formed for 5 min ranged from ~0.13 to ~0.17 during 240 s of the sliding test; a continuous rise was then observed in the friction coefficient after the initial low values, which could indicate substrate removal. During the initial stage of testing, the very low friction coefficient was driven by the modified morphology of the coating. It was found that the PEO coating formed for 10 min showed weaker wear protection under the same load. The large gap (~20 μ m ) between the inner layer and the outer layer caused the weak tribological behavior. The transition point (end of coating life) for the coating formed for 10 min occurred much faster than the coating formed for 5 min. Finally, the friction coefficient of the coating formed for 30 min rose significantly from ~0.4 to ~0.67 during 344 s as the initial stage of testing. After that, the coefficient remained constant until ~1252 s, and then abruptly declined to the lowest value of ~0.47 at 1394 s. Figure 5d displays the profile along the cross-sections of the wear scars after the dry sliding tests for coatings produced with different PEO treatment times. The weakest wear performance with the deepest scar was observed for the PEO coating produced for 10 min due to the early collapse of the compact outer layer, and thus exposing the thin inner layer. The greatest wear performance was recorded for the coating formed for 30 min with a wear depth of ~20 μ m .
The surface cross-section image of the wear scar remained after dry sliding wear test for the PEO coating produced in aluminate electrolyte for 30 min were shown in Figure 6a. It was evident that the whole outer layer was removed, while the inner layer was intact. The outstanding performance of the coating stemmed from the mechanical strength of the outer layer and the great adhesion of the inner layer to the substrate [54,65]. Figure 6b shows the impact of the sliding during wear testing from the central part of the wear scar. Due to extended sliding, an iron oxide layer remained on the wear track. The patches with the transfer layer protruded above the surrounding inner layer and supported the contact points with the steel ball. The remnant debris from the previous round of sliding enhanced the wear resistance during the sliding process and acted as lubrication [65]. The sliding wear tests were also performed on the coatings formed in concentrated aluminate electrolyte [54]. Furthermore, the PEO coating formed for 10 min in concentrated aluminate electrolyte with the thickness of ~64.4 µm was tested under a high load of 30 N and compared with the coating formed in dilute aluminate electrolyte for 30 min with a thickness of ~75.8 µm. Figure 6c,d demonstrate the cross-sectional profiles of wear scars due to the dry sliding test for 30 min for comparing PEO coatings formed in concentrated and dilute aluminate electrolyte and for comparing PEO coatings formed in aluminate and phosphate electrolyte [54]. The low friction coefficient at early times could be related to lack of formation of the transfer layer on the coating surface; however, developing the transfer layer due to frictional heating and mechanical stresses caused the friction coefficient to rise [66]. Cheng et al., finally concluded that the PEO coating formed for 10 min demonstrated greater wear protections, while the wear depth was ~20 µm after 30 min of dry sliding with the applied load of 30 N.
The presence of nanoplate-like α -Al2O3 out-layered Al2O3–ZrO2 composite PEO coating enhanced the wear protection of PEO coating significantly [67,68], highlighting its potential applicability for articular head replacement. It was found that the presence of tetragonal zirconia could enhance the protection performance of the coating produced in aluminate electrolyte, in which the applied load could expedite the transformation of zirconia from tetragonal to monoclinic phases. The volume expansion caused by the transformation induced dilatational and shear stresses, preventing the progression of cracks, and thus improving the wear performance [68].
Wei et al., worked on the wear mechanism of PEO coating produced in phosphate electrolyte [69]. They found that the lowest wear rate of the PEO coating was 1/60 of bare Zirlo alloy in similar conditions, and the wear mechanism of PEO coating was primarily the adhesive wear, where the substrate demonstrated the abrasive wear [69]. Adding graphene oxide (GO) particles in electrolyte also showed a positive effect on reducing friction and improved the fretting wear resistance [8]. The PEO coating with GO addition showed delamination, wear debris, and channel trace, demonstrating abrasive wear as the dominant wear mechanism. However, the PEO coating with GO did not show worn-out protrusion in the surface morphology. Recently, Li et al., conducted a comparative study on the influence of different nanoparticles, including Al2O3, MoS2, CeO2, and graphene oxide (GO), on fretting corrosion behavior [7]. Basically, the nanoparticles embedded in coatings acted as a lubrication phase to decrease the friction. The composite PEO coatings with MoS2 and GO showed low coefficient of friction value in the initial stage of the wear test, after detaching the nanoparticles, the coefficient of friction rapidly increased. It was suggested that abrasive wear was the main wear mechanism for the PEO coating with Al2O3, while the wear mechanism for the PEO coating with either MoS2 or CeO2 was abrasive wear and adhesive wear. For the PEO coating with GO, the wear mechanism was abrasive wear. The PEO composite coating with GO showed that the wear volume was lowest among other composite coatings, but its tribocorrosion performance revealed the highest material loss due to the impact of wear during the corrosion test. The presence of GO particles could effectively reduce the friction and wear as GOs had small shear stress between the layers making the sliding process easier.

5. Photoluminescence Performance

ZrO2 based materials have drawn considerable interest from many researchers in photonics, optoelectronics, and electronics due to their outstanding thermal, mechanical, electrical, and optical properties [70,71,72,73]. Optical transparency in the visible and near infrared region, high refractive index, low optical loss, wide energy bandgap (~5 eV), and low phonon frequency (~470 cm−1) have demonstrated zirconia as a potential matrix host for fabrication of doped compounds using highly efficient luminescent materials ranging from rare-earth to transition metals [74,75,76,77,78,79]. Recently, researchers have intended to take advantage of PEO processing to entrap rare-earth oxide particles in the ZrO2 matrix [74,75,76,77,78,79,80,81].
Figure 7a demonstrated diffuse reflection spectra of the PEO coating produced in citric acid electrolyte for a range of treatment times from 1 to 10 min [82]. A wide absorption band was detected spanning the range of 200 to 330 nm. The main absorption peak occurred at approximately 236 nm (~5.2 eV in photon energy) that could be related to photon excitation from valance band to conduction band [83,84]. In addition, the absorption peak at approximately 294 nm (~4.3 eV in photon energy) could be related to the interstitial Zr3+ ions [85]. Photoluminescence (PL) measurements displayed easily identifiable PL bands ranging from 300 to 600 nm at ambient temperature. The emission spectra divulged 4 peaks centered approximately at 418, 440, 464, and 495 nm. The prolonged PEO treatment time increased the PL intensity because the concentration of oxygen vacancies enhanced in m-ZrO2 films, which was responsible for PL excitation bands. This behavior was also recorded for photocatalytic activity of zirconia films, in which prolonged PEO treatment time enhanced the activity as a result of entrapping more oxygen vacancies in the thicker film [82]. Stojadinović et al., worked on the possibility of fabricating doped ZrO2 coatings with Ho3+ and Ho3+/Yb3+ using the PEO process in phosphate electrolyte containing Ho2O3 or H2O3/Yb2O3 particles [78]. They investigated up- and down-conversion PL properties of the complex oxide coatings composed of ZrO2 as a host and rare-earth oxide as components of the composite. Notably, rare-earth oxide particles featured zeta potentials in the negative range for alkaline environments, and thereby, the applied potential in the PEO process sucked Ho2O3 and Yb2O3 particles toward the anode and entrapped them inside microdischarge channels to form a composite of mixed oxides [78]. The PL emission spectra of PEO coatings evolved with different quantities of Ho2O3 particles and excited with 280 nm radiation, as illustrated in Figure 7b. Pure ZrO2 coating could only display a wide PL band with a spectral maximum approximately at 490 nm in the visible region originating from optical transitions in PL centers derived from oxygen vacancies of ZrO2 [82].
Therefore, two overlapping regions could be found in down-conversion PL spectra of the composite coating excited by 280 nm including the broad PL bond from the matrix and several emission bands related to f–f transition of Ho3+ ions [78]. At around 540 and 550 nm, there were two predominant peaks relating to F 5 4 I 5 8 and S 5 2 I 5 8 transitions, respectively. Increasing treatment time of the PEO processing caused intensifying the wide emission bond of the PL intensity of the ZrO2 matrix and enhancing the emission bonds of Ho3. It was apparent that the PL intensity was greater for un-doped coatings, where higher concentration of Ho3+ ions participating in the ZrO2 matrix caused reducing the PL intensity of the broad emission band, indicating the presence of a non-radiative energy transient between matrix and dopants [78]. The PL excitation spectra were recorded at 540 nm for both doped and un-doped zirconia coatings, as shown in Figure 7c. While the excitation PL spectrum of un-doped zirconia depicted a wide band centered at about 280 nm, the excitation spectrum of Ho-doped zirconia demonstrated an intensified wide band ranging from 250 to 325 nm and several sharp peaks between 325 and 500 nm. Electron transit between fully occupied 2p orbitals of O2− ions of ZrO2 and 4f orbitals of Ho3+ ions led to the formation of the early broadened band, while sharp peaks were related to direct excitation from I 5 8 ground state to the greater energy level of the 4f-manifold. Peaks and assigned electronic transition are clearly illustrated in Figure 7c. Therefore, the PL emission intensity of doped zirconia coatings with rare-earth ions could alter with changing the thickness of the PEO coatings and concentration of rare-earth ions participating in the composite coating [78]. The down-conversion PL emission spectra of PEO coatings formed in electrolyte containing H2O3 and different concentrations of Yb2O3 particles are also shown for excitements at 280 and 450 nm radiations in Figure 7d. The addition of Yb3+ ions into the composite of ZrO2:Ho3+ coatings did not alter the overall shapes of the PL spectra, however, the intensity of the wideband derived from the zirconia matrix and PL bands related to f–f transition of Ho3+ ions were enhanced by raising the incorporation of Yb3+ ions into the composite coatings. This behavior was an indication of the role of Yb3+ ions in transmitting energy to Ho3+ ions [78,86].
Ćirić and Stojadinović also produced the ZrO2 coating with Pr3+ incorporated particles [76]. Accordingly, they found that increasing the treatment time of the PEO process also increased the incorporation of uniformly distributed Pr3+ in the composite coating, where its contribution concurrently stabilized the tetragonal zirconia phase. They categorized PL emission spectral features into two distinct regions, including a broad PL band centered at approximately 495 nm and secondly a region due to the 4f–4f transitions of Pr3+, where the most intense one was from the D 1 2 H 3 4 transition [76]. The same authors reported fabrication of doped composite zirconia coatings containing Tm3+ and Yb3+ using the PEO process [75]. Tm and Yb particles promoted phase transformation from monoclinic to tetragonal phase, where the PL excitation spectra were identified by the broad band of ZrO2 centered at approximately 280 nm and a peak at approximately 359 nm related to the H 3 6 D 1 2 transition of Tm3+.

6. Concluding Remarks and Future Outlook

The PEO coatings have various applications upon the preparation process and zirconium substrate. Several methods have already been implemented to improve bioresponse and antibacterial activity on the Zr surface. Until now, the PEO process has demonstrated promising results in reducing the risks associated with using the PEO coated Zr. Particularly, medical implants demand significant surface biocompatibility, which bare zirconium cannot meet for certain criteria. The rough and porous surface of PEO coatings stimulates nucleation and growth of biocompatible hydroxyapatite compounds due to offering a larger available surface for ionic interactions. The formation of PEO coatings reduces bacterial adhesion, and the incorporation of antibacterial agents, such as silver ions, provides a greater antibacterial feature. Incorporation of Zn and Cu ions in the PEO coating enhances the biocompatibility of coatings, where a more uniform and compact hydroxyapatite forms on the surface of coatings after prolonged immersion in the simulated body fluid.
PEO coating can enhance the corrosion protection of zirconium when immersed in aggressive media. The coatings were able to block the penetration paths of aggressive ions and enhanced the endurance of the zirconium substrate. The PEO coating also reduced the coefficient of friction and offered greater protection against sliding wear, as the debris released from the coatings usually acted as lubrication. Besides, the PEO composite coating with rare-earth elements offered tuning photoluminescence features of the zirconia, where the addition of different rare earth oxides could tailor the excitation and emission range of composite zirconia oxide.
On account of reviewed studies, the following remarks can be suggested for future studies:
  • Inducing new nanoparticles in electrolytes during the PEO process opens a new path for future research, where there is a lack of studies on biomedical applications of Zr and Zr-alloys, such as enhancing the coatings’ bioresponse and efficient delivering drugs to the body tissues.
  • Poor surface features such as numerous pores and cracks confine the protection mechanism to the barrier inhibition, in which the coating halts the penetration of aggressive ions. However, producing composite coatings with protective agents such as sacrificing anodic nanoparticles can endow active protection to the substrates.
  • Wear behavior of coatings requires significant improvements. Besides, the wear mechanism of the coating has not been studied comprehensively. Using the PEO coated Zr-alloys in nuclear industries and applications demanding significant protection against fretting wear and corrosion requires more studies with simulated conditions.
  • Tuning ZrO2 bandgap for photocatalytic applications can draw attention to producing composite coatings with versatile applicability in splitting water and other compounds.

Author Contributions

C.V.R. conceived and supervised this work. N.A. collected all the data, obtained permission to use the data existing in the literature, and compiled the scientific validations. Both authors involved in preparing the manuscript. Both authors have read and agreed to the published version of the manuscript.

Funding

The authors also acknowledge, with pleasure, support from the National Science Foundation (NSF) with NSF-PREM grant #DMR-1827745.

Acknowledgments

NA acknowledges with pleasure the technical support and encouragement provided by the Center for Advanced Materials Research (CMR), UTEP. NA also acknowledges the Research Associate opportunity provided by CMR, UTEP.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fattah-alhosseini, A.; Molaei, M.; Attarzadeh, N.; Babaei, K.; Attarzadeh, F. On the Enhanced Antibacterial Activity of Plasma Electrolytic Oxidation (PEO) Coatings That Incorporate Particles: A Review. Ceram. Int. 2020, 46, 20587–20607. [Google Scholar] [CrossRef]
  2. Fattah-alhosseini, A.; Molaei, M.; Babaei, K. The Effects of Nano- and Micro-Particles on Properties of Plasma Electrolytic Oxidation (PEO) Coatings Applied on Titanium Substrates: A Review. Surf. Interfaces 2020, 21, 100659. [Google Scholar] [CrossRef]
  3. Molaei, M.; Attarzadeh, N.; Fattah-alhosseini, A. Tailoring the Biological Response of Zirconium Implants Using Zirconia Bioceramic Coatings: A Systematic Review. J. Trace Elem. Med. Biol. 2021, 66, 126756. [Google Scholar] [CrossRef]
  4. Bordbar-Khiabani, A.; Yarmand, B.; Sharifi-Asl, S.; Mozafari, M. Improved Corrosion Performance of Biodegradable Magnesium in Simulated Inflammatory Condition via Drug-Loaded Plasma Electrolytic Oxidation Coatings. Mater. Chem. Phys. 2020, 239, 122003. [Google Scholar] [CrossRef]
  5. Wang, D.; Liu, X.; Wu, Y.; Han, H.; Yang, Z.; Su, Y.; Zhang, X.; Wu, G.; Shen, D. Evolution Process of the Plasma Electrolytic Oxidation (PEO) Coating Formed on Aluminum in an Alkaline Sodium Hexametaphosphate ((NaPO3)6) Electrolyte. J. Alloy. Compd. 2019, 798, 129–143. [Google Scholar] [CrossRef]
  6. Sankara Narayanan, T.S.N.; Park, I.S.; Lee, M.H. Strategies to Improve the Corrosion Resistance of Microarc Oxidation (MAO) Coated Magnesium Alloys for Degradable Implants: Prospects and Challenges. Prog. Mater. Sci. 2014, 60, 1–71. [Google Scholar] [CrossRef]
  7. Li, Z.; Cai, Z.; Cui, X.-J.; Liu, R.; Yang, Z.; Zhu, M. Influence of Nanoparticle Additions on Structure and Fretting Corrosion Behavior of Micro-Arc Oxidation Coatings on Zirconium Alloy. Surf. Coat. Technol. 2021, 410, 126949. [Google Scholar] [CrossRef]
  8. Li, Z.; Cai, Z.; Ding, Y.; Cui, X.-J.; Yang, Z.; Zhu, M. Characterization of Graphene Oxide/ZrO2 Composite Coatings Deposited on Zirconium Alloy by Micro-Arc Oxidation. Appl. Surf. Sci. 2020, 506, 144928. [Google Scholar] [CrossRef]
  9. Arun, S.; Arunnellaiappan, T.; Rameshbabu, N. Fabrication of the Nanoparticle Incorporated PEO Coating on Commercially Pure Zirconium and Its Corrosion Resistance. Surf. Coat. Technol. 2016, 305, 264–273. [Google Scholar] [CrossRef]
  10. Qin, W.; Nam, C.; Li, H.L.; Szpunar, J.A. Tetragonal Phase Stability in ZrO2 Film Formed on Zirconium Alloys and Its Effects on Corrosion Resistance. Acta Mater. 2007, 55, 1695–1701. [Google Scholar] [CrossRef]
  11. Cheng, Y.; Matykina, E.; Arrabal, R.; Skeldon, P.; Thompson, G.E. Plasma Electrolytic Oxidation and Corrosion Protection of Zircaloy-4. Surf. Coat. Technol. 2012, 206, 3230–3239. [Google Scholar] [CrossRef]
  12. Sandhyarani, M.; Rameshbabu, N.; Venkateswarlu, K.; Sreekanth, D.; Subrahmanyam, C. Surface Morphology, Corrosion Resistance and in Vitro Bioactivity of P Containing ZrO2 Films Formed on Zr by Plasma Electrolytic Oxidation. J. Alloy. Compd. 2013, 553, 324–332. [Google Scholar] [CrossRef]
  13. Yan, Y.; Han, Y.; Li, D.; Huang, J.; Lian, Q. Effect of NaAlO2 Concentrations on Microstructure and Corrosion Resistance of Al2O3/ZrO2 Coatings Formed on Zirconium by Micro-Arc Oxidation. Appl. Surf. Sci. 2010, 256, 6359–6366. [Google Scholar] [CrossRef]
  14. Gefen, A. Optimizing the Biomechanical Compatibility of Orthopedic Screws for Bone Fracture Fixation. Med. Eng. Phys. 2002, 24, 337–347. [Google Scholar] [CrossRef]
  15. Thomsen, P.; Larsson, C.; Ericson, L.E.; Sennerby, L.; Lausmaa, J.; Kasemo, B. Structure of the Interface between Rabbit Cortical Bone and Implants of Gold, Zirconium and Titanium. J. Mater. Sci. Mater. Med. 1997, 8, 653–665. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Y.-T.; Lee, T.-M.; Lui, T.-S. Enhanced Osteoblastic Cell Response on Zirconia by Bio-Inspired Surface Modification. Colloids Surf. B Biointerfaces 2013, 106, 37–45. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, L.; Hu, X.; Nie, X. Deposition and Properties of Zirconia Coatings on a Zirconium Alloy Produced by Pulsed DC Plasma Electrolytic Oxidation. Surf. Coat. Technol. 2013, 221, 150–157. [Google Scholar] [CrossRef]
  18. Sanchez, A.G.; Schreiner, W.; Duffó, G.; Ceré, S. Surface Characterization of Anodized Zirconium for Biomedical Applications. Appl. Surf. Sci. 2011, 257, 6397–6405. [Google Scholar] [CrossRef]
  19. Zhou, F.Y.; Wang, B.L.; Qiu, K.J.; Li, H.F.; Li, L.; Zheng, Y.F.; Han, Y. In Vitro Corrosion Behavior and Cellular Response of Thermally Oxidized Zr–3Sn Alloy. Appl. Surf. Sci. 2013, 265, 878–888. [Google Scholar] [CrossRef]
  20. Brenier, R.; Mugnier, J.; Mirica, E. XPS Study of Amorphous Zirconium Oxide Films Prepared by Sol–Gel. Appl. Surf. Sci. 1999, 143, 85–91. [Google Scholar] [CrossRef]
  21. Leushake, U.; Krell, T.; Schulz, U.; Peters, M.; Kaysser, W.A.; Rabin, B.H. Microstructure and Phase Stability of EB-PVD Alumina and Alumina/Zirconia for Thermal Barrier Coating Applications. Surf. Coat. Technol. 1997, 94–95, 131–136. [Google Scholar] [CrossRef]
  22. Shanmugavelayutham, G.; Yano, S.; Kobayashi, A. Microstructural Characterization and Properties of ZrO2/Al2O3 Thermal Barrier Coatings by Gas Tunnel-Type Plasma Spraying. Vacuum 2006, 80, 1336–1340. [Google Scholar] [CrossRef]
  23. Zhang, Z.; Wang, K.; Bai, C.; Li, X.; Dang, X.; Zhang, C. The Influence of UV Irradiation on the Biological Properties of MAO-Formed ZrO2. Colloids Surf. B Biointerfaces 2012, 89, 40–47. [Google Scholar] [CrossRef]
  24. Sandhyarani, M.; Prasadrao, T.; Rameshbabu, N. Role of Electrolyte Composition on Structural, Morphological and in-Vitro Biological Properties of Plasma Electrolytic Oxidation Films Formed on Zirconium. Appl. Surf. Sci. 2014, 317, 198–209. [Google Scholar] [CrossRef]
  25. Han, Y.; Yan, Y.; Lu, C. Ultraviolet-Enhanced Bioactivity of ZrO2 Films Prepared by Micro-Arc Oxidation. Thin Solid Film. 2009, 517, 1577–1581. [Google Scholar] [CrossRef]
  26. Han, Y.; Yan, Y.; Lu, C.; Zhang, Y.; Xu, K. Bioactivity and Osteoblast Response of the Micro-Arc Oxidized Zirconia Films. J. Biomed. Mater. Res. Part A 2009, 88A, 117–127. [Google Scholar] [CrossRef] [PubMed]
  27. Yan, Y.; Han, Y.; Lu, C. The Effect of Chemical Treatment on Apatite-Forming Ability of the Macroporous Zirconia Films Formed by Micro-Arc Oxidation. Appl. Surf. Sci. 2008, 254, 4833–4839. [Google Scholar] [CrossRef]
  28. Zhang, L.; Zhu, S.; Han, Y.; Xiao, C.; Tang, W. Formation and Bioactivity of HA Nanorods on Micro-Arc Oxidized Zirconium. Mater. Sci. Eng. C 2014, 43, 86–91. [Google Scholar] [CrossRef] [PubMed]
  29. Cengiz, S.; Uzunoglu, A.; Stanciu, L.; Tarakci, M.; Gencer, Y. Direct Fabrication of Crystalline Hydroxyapatite Coating on Zirconium by Single-Step Plasma Electrolytic Oxidation Process. Surf. Coat. Technol. 2016, 301, 74–79. [Google Scholar] [CrossRef]
  30. Surmenev, R.A.; Surmeneva, M.A.; Ivanova, A.A. Significance of Calcium Phosphate Coatings for the Enhancement of New Bone Osteogenesis—A Review. Acta Biomater. 2014, 10, 557–579. [Google Scholar] [CrossRef] [PubMed]
  31. Viswanath, B.; Ravishankar, N. Controlled Synthesis of Plate-Shaped Hydroxyapatite and Implications for the Morphology of the Apatite Phase in Bone. Biomaterials 2008, 29, 4855–4863. [Google Scholar] [CrossRef]
  32. Roy, M.; Bandyopadhyay, A.; Bose, S. Induction Plasma Sprayed Nano Hydroxyapatite Coatings on Titanium for Orthopaedic and Dental Implants. Surf. Coat. Technol. 2011, 205, 2785–2792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Lu, S.-F.; Lou, B.-S.; Yang, Y.-C.; Wu, P.-S.; Chung, R.-J.; Lee, J.-W. Effects of Duty Cycle and Electrolyte Concentration on the Microstructure and Biocompatibility of Plasma Electrolytic Oxidation Treatment on Zirconium Metal. Thin Solid Film. 2015, 596, 87–93. [Google Scholar] [CrossRef]
  34. Bayati, M.R.; Aminzare, M.; Molaei, R.; Sadrnezhaad, S.K. Micro Arc Oxidation of Nano-Crystalline Ag-Doped TiO2 Semiconductors. Mater. Lett. 2011, 65, 840–842. [Google Scholar] [CrossRef]
  35. Teker, D.; Muhaffel, F.; Menekse, M.; Karaguler, N.G.; Baydogan, M.; Cimenoglu, H. Characteristics of Multi-Layer Coating Formed on Commercially Pure Titanium for Biomedical Applications. Mater. Sci. Eng. C 2015, 48, 579–585. [Google Scholar] [CrossRef]
  36. Fidan, S.; Muhaffel, F.; Riool, M.; Cempura, G.; de Boer, L.; Zaat, S.A.J.; Filemonowicz, A.C.; Cimenoglu, H. Fabrication of Oxide Layer on Zirconium by Micro-Arc Oxidation: Structural and Antimicrobial Characteristics. Mater. Sci. Eng. C 2017, 71, 565–569. [Google Scholar] [CrossRef]
  37. Durdu, S.; Aktug, S.L.; Aktas, S.; Yalcin, E.; Cavusoglu, K.; Altinkok, A.; Usta, M. Characterization and in Vitro Properties of Antibacterial Ag-Based Bioceramic Coatings Formed on Zirconium by Micro Arc Oxidation and Thermal Evaporation. Surf. Coat. Technol. 2017, 331, 107–115. [Google Scholar] [CrossRef]
  38. Aktuğ, S.L.; Durdu, S.; Yalçın, E.; Çavuşoğlu, K.; Usta, M. In Vitro Properties of Bioceramic Coatings Produced on Zirconium by Plasma Electrolytic Oxidation. Surf. Coat. Technol. 2017, 324, 129–139. [Google Scholar] [CrossRef]
  39. Aktug, S.L.; Kutbay, I.; Usta, M. Characterization and Formation of Bioactive Hydroxyapatite Coating on Commercially Pure Zirconium by Micro Arc Oxidation. J. Alloy. Compd. 2017, 695, 998–1004. [Google Scholar] [CrossRef]
  40. Cengiz, S.; Azakli, Y.; Tarakci, M.; Stanciu, L.; Gencer, Y. Microarc Oxidation Discharge Types and Bio Properties of the Coating Synthesized on Zirconium. Mater. Sci. Eng. C 2017, 77, 374–383. [Google Scholar] [CrossRef]
  41. Sandhyarani, M.; Rameshbabu, N.; Venkateswarlu, K. Fabrication, Characterization and in-Vitro Evaluation of Nanostructured Zirconia/Hydroxyapatite Composite Film on Zirconium. Surf. Coat. Technol. 2014, 238, 58–67. [Google Scholar] [CrossRef]
  42. Durdu, S.; Aktug, S.L.; Aktas, S.; Yalcin, E.; Usta, M. Fabrication and in Vitro Properties of Zinc-Based Superhydrophilic Bioceramic Coatings on Zirconium. Surf. Coat. Technol. 2018, 344, 467–478. [Google Scholar] [CrossRef]
  43. Zhang, L.; Han, Y.; Tan, G. Hydroxyaptite Nanorods Patterned ZrO2 Bilayer Coating on Zirconium for the Application of Percutaneous Implants. Colloids Surf. B Biointerfaces 2015, 127, 8–14. [Google Scholar] [CrossRef]
  44. Aktug, S.L.; Durdu, S.; Aktas, S.; Yalcin, E.; Usta, M. Characterization and Investigation of in Vitro Properties of Antibacterial Copper Deposited on Bioactive ZrO2 Coatings on Zirconium. Thin Solid Film. 2019, 681, 69–77. [Google Scholar] [CrossRef]
  45. Cengiz, S.; Uzunoglu, A.; Huang, S.M.; Stanciu, L.; Tarakci, M.; Gencer, Y. An In-Vitro Study: The Effect of Surface Properties on Bioactivity of the Oxide Layer Fabricated on Zr Substrate by PEO. Surf. Interfaces 2021, 22, 100884. [Google Scholar] [CrossRef]
  46. Motta, A.T.; Yilmazbayhan, A.; da Silva, M.J.G.; Comstock, R.J.; Was, G.S.; Busby, J.T.; Gartner, E.; Peng, Q.; Jeong, Y.H.; Park, J.Y. Zirconium Alloys for Supercritical Water Reactor Applications: Challenges and Possibilities. J. Nucl. Mater. 2007, 371, 61–75. [Google Scholar] [CrossRef]
  47. Yau, T.-L.; Annamalai, V.E. Corrosion of Zirconium and its Alloys. In Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2016; ISBN 978-0-12-803581-8. [Google Scholar]
  48. Yau, T.-L.; Sutherlin, R.C.; Chang, A.T.I.W. Corrosion of Zirconium and Zirconium Alloys. In Corrosion: Materials; Stephen, D.C., Bernard, S.C., Eds.; ASM International: Ohio, OH, USA, 2018; Chapter 20; pp. 300–324. [Google Scholar]
  49. Chen, Y.; Nie, X.; Northwood, D.O. Investigation of Plasma Electrolytic Oxidation (PEO) Coatings on a Zr–2.5Nb Alloy Using High Temperature/Pressure Autoclave and Tribological Tests. Surf. Coat. Technol. 2010, 205, 1774–1782. [Google Scholar] [CrossRef]
  50. Cheng, Y.; Wu, F. Plasma Electrolytic Oxidation of Zircaloy-4 Alloy with DC Regime and Properties of Coatings. Trans. Nonferrous Met. Soc. China 2012, 22, 1638–1646. [Google Scholar] [CrossRef]
  51. Cheng, Y.; Wu, F.; Matykina, E.; Skeldon, P.; Thompson, G.E. The Influences of Microdischarge Types and Silicate on the Morphologies and Phase Compositions of Plasma Electrolytic Oxidation Coatings on Zircaloy-2. Corros. Sci. 2012, 59, 307–315. [Google Scholar] [CrossRef]
  52. Sreekanth, D.; Rameshbabu, N.; Venkateswarlu, K. Effect of Various Additives on Morphology and Corrosion Behavior of Ceramic Coatings Developed on AZ31 Magnesium Alloy by Plasma Electrolytic Oxidation. Ceram. Int. 2012, 38, 4607–4615. [Google Scholar] [CrossRef]
  53. Cheng, Y.; Wu, F.; Dong, J.; Wu, X.; Xue, Z.; Matykina, E.; Skeldon, P.; Thompson, G.E. Comparison of Plasma Electrolytic Oxidation of Zirconium Alloy in Silicate- and Aluminate-Based Electrolytes and Wear Properties of the Resulting Coatings. Electrochim. Acta 2012, 85, 25–32. [Google Scholar] [CrossRef]
  54. Cheng, Y.; Cao, J.; Peng, Z.; Wang, Q.; Matykina, E.; Skeldon, P.; Thompson, G.E. Wear-Resistant Coatings Formed on Zircaloy-2 by Plasma Electrolytic Oxidation in Sodium Aluminate Electrolytes. Electrochim. Acta 2014, 116, 453–466. [Google Scholar] [CrossRef]
  55. Trivinho-Strixino, F.; da Silva, D.X.; Paiva-Santos, C.O.; Pereira, E.C. Tetragonal to Monoclinic Phase Transition Observed during Zr Anodisation. J. Solid State Electrochem. 2013, 17, 191–199. [Google Scholar] [CrossRef]
  56. Zou, Z.; Xue, W.; Jia, X.; Du, J.; Wang, R.; Weng, L. Effect of Voltage on Properties of Microarc Oxidation Films Prepared in Phosphate Electrolyte on Zr–1Nb Alloy. Surf. Coat. Technol. 2013, 222, 62–67. [Google Scholar] [CrossRef]
  57. Xue, W.; Zhu, Q.; Jin, Q.; Hua, M. Characterization of Ceramic Coatings Fabricated on Zirconium Alloy by Plasma Electrolytic Oxidation in Silicate Electrolyte. Mater. Chem. Phys. 2010, 120, 656–660. [Google Scholar] [CrossRef]
  58. Sukumaran, A.; Sampatirao, H.; Balasubramanian, R.; Parfenov, E.; Mukaeva, V.; Nagumothu, R. Formation of ZrO2–SiC Composite Coating on Zirconium by Plasma Electrolytic Oxidation in Different Electrolyte Systems Comprising of SiC Nanoparticles. Trans. Inst. Met. 2018, 71, 1699–1713. [Google Scholar] [CrossRef]
  59. Wang, Y.M.; Feng, W.; Xing, Y.R.; Ge, Y.L.; Guo, L.X.; Ouyang, J.H.; Jia, D.C.; Zhou, Y. Degradation and Structure Evolution in Corrosive LiOH Solution of Microarc Oxidation Coated Zircaloy-4 Alloy in Silicate and Phosphate Electrolytes. Appl. Surf. Sci. 2018, 431, 2–12. [Google Scholar] [CrossRef]
  60. Wu, J.; Lu, P.; Dong, L.; Zhao, M.; Li, D.; Xue, W. Combination of Plasma Electrolytic Oxidation and Pulsed Laser Deposition for Preparation of Corrosion-Resisting Composite Film on Zirconium Alloys. Mater. Lett. 2020, 262, 127080. [Google Scholar] [CrossRef]
  61. Yilmazbayhan, A.; Motta, A.T.; Comstock, R.J.; Sabol, G.P.; Lai, B.; Cai, Z. Structure of Zirconium Alloy Oxides Formed in Pure Water Studied with Synchrotron Radiation and Optical Microscopy: Relation to Corrosion Rate. J. Nucl. Mater. 2004, 324, 6–22. [Google Scholar] [CrossRef]
  62. Raj, B.; Mudali, U.K. Materials Development and Corrosion Problems in Nuclear Fuel Reprocessing Plants. Prog. Nucl. Energy 2006, 48, 283–313. [Google Scholar] [CrossRef]
  63. Fisher, N.J.; Weckwerth, M.K.; Grandison, D.A.E.; Cotnam, B.M. Fretting-Wear of Zirconium Alloys. Nucl. Eng. Des. 2002, 213, 79–90. [Google Scholar] [CrossRef] [Green Version]
  64. Helmi Attia, M. On the Fretting Wear Mechanism of Zr-Alloys. Tribol. Int. 2006, 39, 1320–1326. [Google Scholar] [CrossRef]
  65. Cheng, Y.; Xue, Z.; Wang, Q.; Wu, X.-Q.; Matykina, E.; Skeldon, P.; Thompson, G.E. New Findings on Properties of Plasma Electrolytic Oxidation Coatings from Study of an Al–Cu–Li Alloy. Electrochim. Acta 2013, 107, 358–378. [Google Scholar] [CrossRef]
  66. Martini, C.; Ceschini, L.; Tarterini, F.; Paillard, J.M.; Curran, J.A. PEO Layers Obtained from Mixed Aluminate–Phosphate Baths on Ti–6Al–4V: Dry Sliding Behaviour and Influence of a PTFE Topcoat. Wear 2010, 269, 747–756. [Google Scholar] [CrossRef]
  67. Yan, Y.; Han, Y.; Huang, J. Formation of Al2O3–ZrO2 Composite Coating on Zirconium by Micro-Arc Oxidation. Scr. Mater. 2008, 59, 203–206. [Google Scholar] [CrossRef]
  68. Zhang, L.; Zhang, W.; Han, Y.; Tang, W. A Nanoplate-like α-Al2O3 out-Layered Al2O3-ZrO2 Coating Fabricated by Micro-Arc Oxidation for Hip Joint Prosthesis. Appl. Surf. Sci. 2016, 361, 141–149. [Google Scholar] [CrossRef]
  69. Wei, K.; Chen, L.; Qu, Y.; Yu, J.; Jin, X.; Du, J.; Xue, W.; Zhang, J. Tribological Properties of Microarc Oxidation Coatings on Zirlo Alloy. Surf. Eng. 2019, 35, 692–700. [Google Scholar] [CrossRef]
  70. Atuchin, V.V.; Aliev, V.S.; Kruchinin, V.N.; Ramana, C.V. Optical Properties of ZrO2 Films Fabricated by Ion Beam Sputtering Deposition at Low Temperature. In Proceedings of the 2007 International Forum on Strategic Technology, Ulaanbaatar, Mongolia, 3–6 October 2007; pp. 529–531. [Google Scholar]
  71. Ramana, C.V.; Utsunomiya, S.; Ewing, R.C.; Becker, U.; Atuchin, V.V.; Aliev, V.S.; Kruchinin, V.N. Spectroscopic Ellipsometry Characterization of the Optical Properties and Thermal Stability of ZrO2 Films Made by Ion-Beam Assisted Deposition. Appl. Phys. Lett. 2008, 92, 11917. [Google Scholar] [CrossRef]
  72. Ramana, C.V.; Vemuri, R.S.; Fernandez, I.; Campbell, A.L. Size-Effects on the Optical Properties of Zirconium Oxide Thin Films. Appl. Phys. Lett. 2009, 95, 231905. [Google Scholar] [CrossRef]
  73. Noor-A-Alam, M.; Ramana, C.; Choudhuri, A. Analysis of microstructure and thermal stability of hafnia-zirconia based thermal barrier coatings. In Proceedings of the 9th Annual International Energy Conversion Engineering Conference, International Energy Conversion Engineering Conference (IECEC), San Diego, CA, USA, 31 July–3 August 2011. [Google Scholar]
  74. Lokesha, H.S.; Nagabhushana, K.R.; Chithambo, M.L.; Singh, F. Down and Upconversion Photoluminescence of ZrO2:Er3+ Phosphor Irradiated with 120 MeV Gold Ions. Mater. Res. Express 2020, 7, 64006. [Google Scholar] [CrossRef]
  75. Ćirić, A.; Stojadinović, S. Photoluminescence Studies of ZrO2:Tm3+/Yb3+ Coatings Formed by Plasma Electrolytic Oxidation. J. Lumin. 2019, 214, 116568. [Google Scholar] [CrossRef]
  76. Ćirić, A.; Stojadinović, S. Photoluminescence Properties of Pr3+ Doped ZrO2 Formed by Plasma Electrolytic Oxidation. J. Alloy. Compd. 2019, 803, 126–134. [Google Scholar] [CrossRef]
  77. Stojadinović, S.; Tadić, N.; Vasilić, R. Photoluminescence Properties of Er3+/Yb3+ Doped ZrO2 Coatings Formed by Plasma Electrolytic Oxidation. J. Lumin. 2019, 208, 296–301. [Google Scholar] [CrossRef]
  78. Stojadinović, S.; Tadić, N.; Vasilić, R. Down- and up-Conversion Photoluminescence of ZrO2:Ho3+ and ZrO2:Ho3+/Yb3+ Coatings Formed by Plasma Electrolytic Oxidation. J. Alloy. Compd. 2019, 785, 1222–1232. [Google Scholar] [CrossRef]
  79. Stojadinović, S.; Tadić, N.; Vasilić, R. Down-Conversion Photoluminescence of ZrO2:Er3+ Coatings Formed by Plasma Electrolytic Oxidation. Mater. Lett. 2018, 219, 251–255. [Google Scholar] [CrossRef]
  80. Stojadinović, S.; Tadić, N.; Vasilić, R. Structural and Photoluminescent Properties of ZrO2:Tb3+ Coatings Formed by Plasma Electrolytic Oxidation. J. Lumin. 2018, 197, 83–89. [Google Scholar] [CrossRef]
  81. Stojadinović, S.; Tadić, N.; Vasilić, R. Photoluminescence of Sm3+ Doped ZrO2 Coatings Formed by Plasma Electrolytic Oxidation of Zirconium. Mater. Lett. 2016, 164, 329–332. [Google Scholar] [CrossRef]
  82. Stojadinović, S.; Vasilić, R.; Radić, N.; Grbić, B. Zirconia Films Formed by Plasma Electrolytic Oxidation: Photoluminescent and Photocatalytic Properties. Opt. Mater. 2015, 40, 20–25. [Google Scholar] [CrossRef]
  83. Cao, H.; Qiu, X.; Luo, B.; Liang, Y.; Zhang, Y.; Tan, R.; Zhao, M.; Zhu, Q. Synthesis and Room-Temperature Ultraviolet Photoluminescence Properties of Zirconia Nanowires. Adv. Funct. Mater. 2004, 14, 243–246. [Google Scholar] [CrossRef]
  84. Cong, Y.; Li, B.; Yue, S.; Fan, D.; Wang, X.J. Effect of Oxygen Vacancy on Phase Transition and Photoluminescence Properties of Nanocrystalline Zirconia Synthesized by the One Pot Reaction. J. Phys. Chem. C 2009, 113, 13974–13978. [Google Scholar] [CrossRef]
  85. Kumari, L.; Li, W.Z.; Xu, J.M.; Leblanc, R.M.; Wang, D.Z.; Li, Y.; Guo, H.; Zhang, J. Controlled Hydrothermal Synthesis of Zirconium Oxide Nanostructures and Their Optical Properties. Cryst. Growth Des. 2009, 9, 3874–3880. [Google Scholar] [CrossRef]
  86. Jain, N.; Singh, R.K.; Sinha, S.; Singh, R.A.; Singh, J. Color Tunable Emission through Energy Transfer from Yb3+ Co-Doped SrSnO3:Ho3+ Perovskite Nano-Phosphor. Appl. Nanosci. 2018, 8, 1267–1278. [Google Scholar] [CrossRef]
Figure 1. (a) SEM micrographs of the surface of the PEO coating immersed in SBF for 14 days. Formation of apatite particles is the indication of biocompatibility [24] (Reproduced with permission number: 5043530687884, Elsevier); (b) schematic representation of the PEO formation containing HA top layer [29] (Reproduced with permission number: 5061321012937, Elsevier).
Figure 1. (a) SEM micrographs of the surface of the PEO coating immersed in SBF for 14 days. Formation of apatite particles is the indication of biocompatibility [24] (Reproduced with permission number: 5043530687884, Elsevier); (b) schematic representation of the PEO formation containing HA top layer [29] (Reproduced with permission number: 5061321012937, Elsevier).
Coatings 11 00620 g001
Figure 2. (a,b) The contact angle measurements: bare substrate and the PEO coating are displayed, respectively [29] (Reproduced with permission number: 5061321012937, Elsevier); (c) the plot for illustration of Gram-positive and Gram-negative bacteria’s adhesion on the surface of PEO and Ag-based PEO coatings from [37] (Reproduced with permission number: 5061400310129, Elsevier).
Figure 2. (a,b) The contact angle measurements: bare substrate and the PEO coating are displayed, respectively [29] (Reproduced with permission number: 5061321012937, Elsevier); (c) the plot for illustration of Gram-positive and Gram-negative bacteria’s adhesion on the surface of PEO and Ag-based PEO coatings from [37] (Reproduced with permission number: 5061400310129, Elsevier).
Coatings 11 00620 g002
Figure 3. (a) PDP plot of Zr substrate, simple PEO coating (P), composite coating of PEO and each of Al2O3 (PA), Ce2O3 (PC), and ZrO2 (PZ) [9] (Reproduced with permission number: 5043540009057, Elsevier); (b) PDP plot of bare zirconium, phosphate (P) and silicate (Si) PEO coatings [59] (Reproduced with permission number: 5043540272527, Elsevier).
Figure 3. (a) PDP plot of Zr substrate, simple PEO coating (P), composite coating of PEO and each of Al2O3 (PA), Ce2O3 (PC), and ZrO2 (PZ) [9] (Reproduced with permission number: 5043540009057, Elsevier); (b) PDP plot of bare zirconium, phosphate (P) and silicate (Si) PEO coatings [59] (Reproduced with permission number: 5043540272527, Elsevier).
Coatings 11 00620 g003
Figure 4. (a,b) Nyquist plots of EIS results of phosphate and silicate PEO coating for prolonged immersions, respectively [59]; (c,d) Equivalent circuit and physical model for phosphate and silicate PEO coatings after 50–100 days immersion, respectively [59] (Reproduced with permission number: 5043540272527, Elsevier).
Figure 4. (a,b) Nyquist plots of EIS results of phosphate and silicate PEO coating for prolonged immersions, respectively [59]; (c,d) Equivalent circuit and physical model for phosphate and silicate PEO coatings after 50–100 days immersion, respectively [59] (Reproduced with permission number: 5043540272527, Elsevier).
Coatings 11 00620 g004
Figure 5. (a,b) The sliding impact on surface morphologies for the silicate and aluminate PEO coating, respectively; (c) cross-section image of the sliding scar at the center of the PEO coating formed in aluminate electrolyte for 600 s [53] (Reproduced with permission number: 5043480179625, Elsevier); (d) wear scars profile after dry sliding under 10 N load on PEO coatings formed in electrolyte mixture of aluminate and KOH for various treatment times: a. 5 min, 30 min sliding time; b. 10 min, 923 s sliding time; c. 30 min, 30 min sliding time [54] (Reproduced with permission number: 5043510569542, Elsevier).
Figure 5. (a,b) The sliding impact on surface morphologies for the silicate and aluminate PEO coating, respectively; (c) cross-section image of the sliding scar at the center of the PEO coating formed in aluminate electrolyte for 600 s [53] (Reproduced with permission number: 5043480179625, Elsevier); (d) wear scars profile after dry sliding under 10 N load on PEO coatings formed in electrolyte mixture of aluminate and KOH for various treatment times: a. 5 min, 30 min sliding time; b. 10 min, 923 s sliding time; c. 30 min, 30 min sliding time [54] (Reproduced with permission number: 5043510569542, Elsevier).
Coatings 11 00620 g005
Figure 6. (a) The cross-section image of sliding bar impact on the coating surface formed in aluminate electrolyte [54]; (b) the central part of the wear scar on the PEO coating formed in aluminate electrolyte [54]; (c) the wear profile of the sliding test for PEO coatings produced in dilute and concentrated aluminate electrolyte under 30 N applied load [54]; (d) the wear profile of the sliding test for the PEO coating formed in concentrated aluminate electrolyte (red color) and the PEO coating formed in phosphate electrolyte (green color) under 10 N applied load [54] (Reproduced with permission number: 5046821419713, Elsevier).
Figure 6. (a) The cross-section image of sliding bar impact on the coating surface formed in aluminate electrolyte [54]; (b) the central part of the wear scar on the PEO coating formed in aluminate electrolyte [54]; (c) the wear profile of the sliding test for PEO coatings produced in dilute and concentrated aluminate electrolyte under 30 N applied load [54]; (d) the wear profile of the sliding test for the PEO coating formed in concentrated aluminate electrolyte (red color) and the PEO coating formed in phosphate electrolyte (green color) under 10 N applied load [54] (Reproduced with permission number: 5046821419713, Elsevier).
Coatings 11 00620 g006
Figure 7. (a) Alteration of diffuse reflection spectra of zirconia films based on treatment time of the PEO process [82] (Reproduced with permission number: 5043541253698, Elsevier); (b) photoluminescence emission spectra excited with 280 nm radiation of PEO coating produced in phosphate electrolyte with addition of Ho2O3 powder in different concentrations for a treatment time of 10 min [78]; (c) photoluminescence excitation spectra of undoped and Ho-doped ZrO2 coatings produced by PEO for a treatment time of 10 min in phosphate electrolyte with and without 0.5 g/L Ho2O3 addition [78]; (d) photoluminescence emission spectra excited with 280 and 450 nm radiation of PEO coatings produced in phosphate electrolyte with addition of 4.0 g/L Ho2O3 particles and Yb2O3 particles in different concentrations for a treatment time of 10 min [78] (Reproduced with permission number: 5043550012051, Elsevier).
Figure 7. (a) Alteration of diffuse reflection spectra of zirconia films based on treatment time of the PEO process [82] (Reproduced with permission number: 5043541253698, Elsevier); (b) photoluminescence emission spectra excited with 280 nm radiation of PEO coating produced in phosphate electrolyte with addition of Ho2O3 powder in different concentrations for a treatment time of 10 min [78]; (c) photoluminescence excitation spectra of undoped and Ho-doped ZrO2 coatings produced by PEO for a treatment time of 10 min in phosphate electrolyte with and without 0.5 g/L Ho2O3 addition [78]; (d) photoluminescence emission spectra excited with 280 and 450 nm radiation of PEO coatings produced in phosphate electrolyte with addition of 4.0 g/L Ho2O3 particles and Yb2O3 particles in different concentrations for a treatment time of 10 min [78] (Reproduced with permission number: 5043550012051, Elsevier).
Coatings 11 00620 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Attarzadeh, N.; Ramana, C.V. Plasma Electrolytic Oxidation Ceramic Coatings on Zirconium (Zr) and Zr-Alloys: Part-II: Properties and Applications. Coatings 2021, 11, 620. https://doi.org/10.3390/coatings11060620

AMA Style

Attarzadeh N, Ramana CV. Plasma Electrolytic Oxidation Ceramic Coatings on Zirconium (Zr) and Zr-Alloys: Part-II: Properties and Applications. Coatings. 2021; 11(6):620. https://doi.org/10.3390/coatings11060620

Chicago/Turabian Style

Attarzadeh, Navid, and C. V. Ramana. 2021. "Plasma Electrolytic Oxidation Ceramic Coatings on Zirconium (Zr) and Zr-Alloys: Part-II: Properties and Applications" Coatings 11, no. 6: 620. https://doi.org/10.3390/coatings11060620

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