Home Physical Sciences Tribocorrosion in biomaterials and control techniques: a review
Article Publicly Available

Tribocorrosion in biomaterials and control techniques: a review

  • Umanath Puthillam

    Umanath Puthillam is pursuing his doctoral degree in mechanical engineering at Vellore Institute of Technology (VIT), Vellore. He has a Master’s degree in production engineering and a Bachelor’s degree in automobile engineering. He has 6 years of experience in teaching at graduate level and authored 3 articles in Q1/Q2 journals.

    ORCID logo
    and Renold Elsen Selvam

    Renold Elsen Selvam has published 41 papers in indexed journals, 4 book chapters, and edited 2 conference proceedings. He has also registered 8 patents, 2 of which have been granted. He has secured and completed multiple projects and consultancy work from national and international funding agencies, amounting to 1.5 crores.

    EMAIL logo
Published/Copyright: November 13, 2023

Abstract

Tribocorrosion is getting more and more popular in biomaterials research. The synergism between wear and corrosion is creating deviations from the expected real-world results from individual corrosion or wear studies. The host body consisting of immune system and dissolved proteins makes them highly corrosive which makes the material selection a unique and challenging process for implant materials. The synergism between corrosion and wear leads to shorter implant life. The research on tribocorrosion has bought an insight into this phenomenon and presented ideas to arrest the premature failure of implants. This review focuses on the recent developments in tribocorrosion research and the effectiveness of remedial actions suggested by them. The influence of materials, processing methods and post-processing treatments are also reviewed in detail.

1 Introduction

Biomaterials are defined as those materials used for artificial transplantation to replace a part or function of the body. The procedure should be viable physiologically and should not raise any safety threats to the host and should be affordable (Ratner et al. 2013). Many materials like stainless steel, ultra-high molecular weight poly ethylene (UHMWPE), zirconia, titanium and its alloys etc., can be implanted into the body and can perform the functions of replaced organ reasonably well. Metals are commonly used as a biomaterial for their electrical, mechanical and thermal properties. There are certain criteria that the material should fulfil to be used as a biomaterial. They are,

  1. Non-toxic: a biomaterial should not be toxic unless it is specifically engineered for such requirements.

  2. Biocompatibility: the material should be biocompatible which should facilitate osseointegration and should not cause any trouble to the host leading to rejection.

  3. Physical properties: it is a property dependent on the part to be replaced. The articular cup of the hip joint should have high lubricity whereas the intraocular lens demands suitable optical properties like refraction index.

  4. Mechanical performance: it also depends on the part to be replaced. For example, if we need to replace a heart valve leaflet, the material should be tough and flexible whereas the material used for hip transplantation must be strong and rigid.

  5. Mechanical durability: it denotes the minimum duration that a transplant must perform its function satisfactorily. it also depends on the material to be replaced, a catheter needs to function for 3 days only and a hip joint must perform its duty for at least 10 years.

  6. Surface properties: the ability of the implant to interact with the surroundings depends on the surface properties like roughness, porosity etc. The implant loosening due to micromotion and formation of fibrous tissue indicating the poor interaction between the implant and host tissues is attributed to poor surface chemistry and topography. The properly engineered implant surface can achieve good osseointegration for the implant; that is direct integration with the bone without any fibrous tissue formation (Nasab and Hacssan 2009).

  7. High corrosion and wear resistance: this is to prevent the accumulation of debris in the blood and surroundings of the implant which could produce harmful effects on the human body locally or on the whole body (Manivasagam et al. 2009).

  8. Out of these 7 factors, corrosion and wear resistance is the area which attracts many researchers since it has numerous opportunities to improve. Other material properties mentioned above are limited by their physical nature and difficult to alter the natural state. Moreover, wear and corrosion resistance can be improved with some tweaks at the surface of implant materials. This review paper is focusing on the phenomenon of tribocorrosion and the recent developments in controlling tribo corrosion.

2 Wear of biomaterials

Wear is the progressive loss of material which is produced by the relative motion of opposite surfaces and causes surface damage and eventually results in the failure of the material to perform the desired function. Many factors influence the wear rate of materials such as the hardness and strength of the material apart from the surrounding conditions such as pressure, temperature, lubrication, corrosion etc. which is crucial for biomaterials. Following are the common wear mechanisms found on solid surfaces.

  1. Adhesive wear: if the wear occurs when two or more solid surfaces slide over each other under pressure, it is an adhesive type of wear. Any small projections or asperities on the surface are plastically deformed and then welded together by the high local pressure caused by the sliding. This type of wear can be reduced by using metals with high hardness and large elastic modulus (Stachowiak and Batchelor 2001)

  2. Abrasive wear: whenever a solid object is loaded against a material that has an equal or greater hardness, abrasive wear will take place (Williams 2006).

  3. Corrosive wear: there are situations where a film material layer may be formed by the chemical attack of any one of the contacting bodies. This layer can either inhibit the corrosion by forming a durable lubricating film or accelerate the damage by forming a weak layer which will disintegrate under sliding contact pressure. The wearing down of film can also cause galvanic coupling between the remaining film and the underlying substrate body and result in a high rate of corrosion (Totten and Liang 2004).

  4. Fatigue wear: the repetitive stress under sliding or rolling contact can result in fatigue wear. The impact of this wear is not severe and they can be identified with the help of mechanics of crack initiation, crack growth and fracture (Williams 2006).

  5. Studies show that the major cause of wear in materials is a combination of these mechanisms rather than caused by any one of them. These types of wear occurring in materials are classified as

  6. Oxidative wear: it is a type of adhesive wear where metals exposed to the atmosphere reacts with oxygen and form metal oxides on the surface. This oxide layer protects them from further adhesive wear but gets ruptured by sliding contact and results in wear debris.

  7. Fretting wear: this is also called fretting corrosion, and this is a combination of adhesive, corrosive and abrasive wear. This type of wear is the result of the oscillatory tangential micromotion of contact surfaces. This is a dangerous type of wear as the debris may get converted to abrasive oxides.

  8. Erosive wear: this type of wear is similar to abrasive wear, and it is caused by protruded sharp edges or sharp particles on the counter body. The mechanism differs from the abrasive wear as it produces quick material removal and the magnitude of damage will be higher.

  9. Cavitation: this type of wear will take place even if there are no contact surfaces with the help of liquids and gases. The relative motion between the fluids and the metal produces local pressure variations which could result in the boiling of fluids. When the pressure is restored, the bubbles formed due to boiling ruptures and cause pressure fluctuation which could produce wear (Kato 2002).

3 Corrosion of biomaterials

Corrosion is an electrochemical reaction which depends on the nature of electrolyte reactivity and the surface conditions of the material. A metal act as an anion in the electrochemical cell releases an electron to the surroundings and becomes a metal ion irrespective of the electrolyte. The released electron reacts with other elements based on the electrolyte.

The biological aqueous solution is making it easier for implant material to corrode quickly (Whitesides and Wong 2006). The following reactions will take place inside the body.

(1)MM(n+)+n(electrons)

where M is the implant metal. The mixed potential theory of corrosion states that there should be no net accumulation of charges in corrosion cells. That is, all the electrons released should be consumed by a counterpart reaction. Since the human body is a complex system with plenty of dissolved components, many possible corrosion mechanisms are feasible inside the body (Gilbert and Mali 2012). They are,

  1. General corrosion: it is also known as uniform corrosion where the damage is spread equally over the surface and leaves behind scale or deposits.

  2. Pitting corrosion: here, the implant is attacked at specific or selected locations. The infected area will spread at a rapid rate than normal, and the implant become defective sooner than expected.

  3. Galvanic corrosion: here, the more active metal gets corroded quickly if 2 dissimilar metals are connected electrically and exposed to any electrolyte.

  4. Crevice corrosion: this type of corrosion occurs at narrow openings or spaces between two metal surfaces or between metal and non-metal surfaces.

  5. Fretting corrosion: the relative sliding motion between implant surfaces ruptures the passive film which initiates rapid corrosion of the material. This is called fretting corrosion which produces metal particles of 0.1–1 µm and it is mostly found in metal on metal joints (Fontana 1980).

  6. Intergranular corrosion: this is a corrosion mechanism similar to galvanic corrosion. The difference is that the source of corrosion is the impurities and inclusions present in the implant alloy (Ong et al. 2017), improper heat treatment, welding etc. could lead to this type of corrosion.

  7. Leaching: this type of corrosion is found in alloys with more than 1 phase which is common in biomaterials. The corrosion induced by differences in the grains rather than the grain boundaries is known as leaching (Ong et al. 2017).

  8. Stress corrosion cracking: this is a type of failure in biomaterial implants which is driven by corrosion and related issues. It is common in implants surrounded by chloride ions which are subjected to residual tensile stress as a result of corrosion and fail at a much lower stress level than normal (Raja and Shoji 2007). However, this is a rare phenomenon and is seldom reported.

The surroundings of an implant material are unique and consist of many components which could accelerate the corrosion rate such as chlorine, proteins etc. Hence the experimental results obtained from the standard corrosion tests cannot be used as a reference for biomaterials. The fluids surrounding the biomaterials in the body are highly corrosive and could damage the implants prematurely (Mathew et al. 2009). Corrosion can significantly reduce the fatigue life and ultimate strength of materials. Corroded products also produce local pain and inflammation in the implant region. Tiantian Hui found that the cell viability of Ti6Al4V was reduced from 99.1 % to 0.5 % when the potential from fretting corrosion dropped by a value of −1 V against Ag/AgCl electrode (Hui et al. 2014).

F. Contu et al. found that more than two-thirds of the stainless steel implants are failing due to pitting and crevice corrosion (Contu et al. 2005). Electrochemical properties are analysed to learn about the impact of corrosion and biocompatibility of implant materials. Researchers are studying total joint replacements (TJR) by monitoring metal levels in the blood (De Micheli and Riesgo 1982; Kuhn 1981; Rossi et al. 2000). Major wear metal on metal (MOM) joint is occurring due to sliding contact. Hence pin-on-disk or ball-on-disk experimental setup is used to simulate the real-life scenario, even if it is not perfect due to the absence of body fluids (Saikko 1998; Wright et al. 1982).

Corrosion is an important parameter in the selection and design of implant material. Corrosion can determine the life of an implant and more importantly, it can release toxic materials to the surrounding body parts. These released materials may get deposited in faraway organs too when blood act as the transport mechanism and can even cause cancer in body parts and organs apart from the loosening (Blackwood 2003; Eliaz and Hakshur 2012; Manivasagam et al. 2010; Virtanen 2008). It is common to notice changes from the normal condition in the vicinity of a new implant like inflammation. The pH value in this region could fall as low as 4 and continue the state for several weeks. This acidic nature could initiate or accelerate the corrosion of biomaterials. Several researchers identified the possibility of this corrosion due to the acidic environment (Ciolac et al. 2000; Laing 1973) Apart from the pH value, dissolved oxygen, dissolved bicarbonate, phosphates, cholesterols and phospholipids and chlorides are characteristic defining features of the body fluid. These body fluids are not as aggressive as seawater in case of corrosion. Hence it is logical to believe that these particles do not influence the corrosion rate significantly. Hence the in vitro studies of corrosion are conducted usually in Ring’s solution, Hank’s solution or saline water only. Ring’s solution and Hank’s solution contain bicarbonate and calcium chloride and they are added on the assumption that these components may play a role in the corrosion. Another reason we are not using blood for the studies is coagulation. Some research studies used blood by adding sodium citrate as an anticoagulant. This sodium citrate can influence the corrosion rate by aiding the passivation of Co–Cr–Mo alloys and stainless steels. Hence it is better not to use blood and anticoagulants for the studies since it can skew the final output of the experiment (Pound 2014). Phosphate buffered saline (PBS) solution is the most common fluid used in in vitro experiments because of its ability to maintain the pH constant throughout the experiment (Corbett 2003) although it does not recreate the original scenario due to the absence of minor elements like sulphur contained in amino acids which can cause crevice corrosion in stainless steels (Traisnel et al. 1990).

Although the research on the effect of protein present in the body fluids in corrosion has not produced a definite conclusion yet, they do have an impact on it. This could be due to the non-uniformity of test conditions used in the research process where some people used single protein albumin which is present in abundant quantity in the blood whereas other teams used serums consisting of multiple proteins (Virtanen 2012). Proteins can carry the metal ions far away from the implant surface which can destabilize the electrical stability of the implant metal and initiate or accelerate the metal dissolution (Ratner et al. 2004). Some studies proved that it can change the electrode potential by varying the electron-carrying capacity (Williams 1986). There are some other studies which produce contradictory results that protein can reduce the rate of corrosion when the adsorbed protein layer act as a barrier between the metal and surrounding (Zierold 1924). The oily nature of proteins also makes them act like a lubricant in mating parts of implants and inhibit the wear, thus reducing wear-accelerated corrosion.

The synergism between albumin and H2O2 in corrosion and tribocorrosion is another area of focus in the study of implant materials. H2O2 is formed in an attempt of the body to react to the implant insertion leading to inflammation. Pan et al. found that the presence of H2O2 considerably reduced the corrosion resistance of implant material (Pan et al. 1996). Similar results were obtained in an experiment conducted by Al-Mobarak et al. on Titanium and its alloys used in dental applications which proved that H2O2 reduces the corrosion resistance of Implant materials (Al-Mobarak et al. 2006). Fei Yu et al. studied the synergistic effect of albumin and hydrogen peroxide in accelerating the corrosion of Ti6Al4V for the first time using electrochemical methods. Their independent observations on the effect of H2O2 and albumin agreed with previous results that the former facilitates corrosion while the latter inhibits it. They also found that the combined solution produced more corrosion than individual solutions even though the OCP values of the combined solution were similar to physiological saline in the absence of H2O2 and albumin since they traced more concentrations of Ti, Al and V in the peri-implant region (Yu et al. 2015). The possible explanation for this contradictory result was explained as the suppression of cathodic reaction with the addition of albumin which takes the potential of the material into active regions. Zhang et al. went on to test the time dependency of corrosion behaviour with albumin and hydrogen peroxide. They concluded that the inhibition caused by albumin is temporary (less than 22 h) and long exposure in this environment increases the corrosion rate. They attributed this behaviour to the thinner oxide layer at the implant surface due to the dissolution of peroxide corrosion products in the presence of albumin (Zhang et al. 2018). None of these studies analysed the influence of mechanical loading or the tribocorrosion scenario of the implant material. Gopal et al. conducted a study on fretting tribocorrosion of Ti6Al4V in the presence of H2O2 alone and H2O2 and bovine serum albumin (BSA – act as a source of albumin). The wear scar and wear volume loss of the implant were lowest in the presence of H2O2 (H2O2 +PBS and H2O2 + BSA + PBS) when compared with BSA + PBS and PBS alone. Solution with the highest percentage of H2O2 with BSA produced fewer grooves and larger oxide patches indicating reduced wear loss. They concluded that the upward shift in the open circuit potential during the fretting motion in the presence of H2O2 reduced the coefficient of friction which eventually resulted in fewer loss of material (Gopal and Manivasagam 2020).

It is well known that the output received from an in vitro does not exactly represent the in vivo condition. There are many reasons for this such as variations in pressure and temperature and the presence of serums and proteins present in the body fluids. These elements will cause corrosion to implant material and corrosion is inevitable. Hence, rather than focusing on creating corrosion-free materials, the focus should be on improving corrosion resistance (Gilbert and Mali 2012). Kuhn et al. found the following points regarding in vitro and in vivo comparison.

  1. Organic species like a serum, which accelerates corrosion, should be included in the in vitro test environment created for a more reliable result.

  2. It should be ensured that the presence of elements added to recreate the in vivo atmosphere should not interfere with the electrochemical set-up and skew the results.

  3. The duration of test time is also of great importance. Some implant materials have a very long passivation period. Hence short-term experiments are not suited for such elements. It is better to conduct the test for a minimum of 1000 h.

  4. The selection of the correct specimen and experimental set-up for the test conditions play an important role in reliability. These 2 factors should be able to reproduce the in vivo condition accurately.

  5. A soft or hard tissue formed around the implant material in vivo can suppress corrosion. The lab test condition should find a way to reproduce this tissue scenario for better results (Kuhn et al. 2009).

Biocompatibility is another factor considered for material selection. This property can influence the corrosion rate of a biomaterial. implant size, shape, material composition, surface wettability, surface roughness and charge are some of the factors that influence the biocompatibility of a material (Eliaz 2019). The corrosive property also influences the biocompatibility of a material. The material detached from the implant material may get transported to the surrounding organs and sometimes to the faraway organs. These micro-sized particles can cause adverse effects on those organs.

No biomaterial is completely inert, and inertness varies from material to material. Following are the commonly found impacts of material corrosion inside the body.

  1. Cell behaviour is affected by electric current.

  2. Changes in chemical composition in the surroundings.

  3. Variations in cellular metabolism (Bhat 2002; Mears 1979).

The metal ions are released because of corrosion to the blood in the case of metallic biomaterials. These ions can cause many effects on the body. It can cause inflammation or pain in the implanted area. Some metal ions could disturb the total balance of constituents in the body and hence the physiological tolerance to toxicity (Bhat 2002; Murphy et al. 2016) as the metal ions floating in the body will combine with biomolecules and become toxic. The level of toxicity depends on the toxicity of the metal ion itself (Hanawa 2012). This change could be visible in the vicinity or organs situated far away from the implant.

3.1 Corrosion of additive manufactured biomaterials and effect of process parameters

3-D printing technology is quickly being adapted for manufacturing in the biomedical industry since it has many advantages and patient-specific designing is one of the most important ones among them (Morris et al. 2021; Manivasagam et al. 2010). Porous implants are proven to be better than cast alloys due to their porosity which can imitate bones better (Alvarez and Nakajima 2009). Selective laser melting (SLM) and electron beam melting (EBM) are the two technologies used to 3D print metallic biomaterials (Heinl et al. 2008). It has been proved that SBM Ti6Al4V produced using the EBM technique is non-cytotoxic (Tuomi et al. 2017). Apart from all the mechanical and biological advantages of Ti6Al4V over the other implant materials, they have better corrosion resistance too due to the formation of a passive oxide layer on the substrate surface (Keselowsky et al. 2007; Kumari et al. 2010). However, this oxide layer of nanometre-level thickness does not have strong mechanical wear resistance and causes damage to the implant in tribocorrosion conditions. Studies show that porosity level can influence the corrosion rate, the higher the porosity, the lower will be the corrosion (Morris et al. 2021). Revilla, et al. found out that aluminium alloy (AlSi10Mg) produced using selective laser melting showed better corrosion resistance than the casted counterpart of the same as Si–Mg segregation was absent in SLM product as in cast product which caused localized corrosion due to galvanic coupling (Revilla et al. 2017). Another team analysed the influence of surface finish on AlSi10Mg produced using the SLM method and found that samples polished using 1200 grit paper always produced a lower corrosion rate than unpolished workpieces (Leon and Aghion 2017). Titanium alloys are the most popular metal used for implants today. Many studies have already been conducted on the corrosion behaviour of titanium and its alloys. Heat treatment is an effective method to improve the mechanical properties of metals and alloys. However, Dai et al. found that heat treatment on Ti6Al4V reduced the corrosion resistance and this reduction was a function of the temperature used in the process (Dai et al. 2017). This was confirmed by Chandramohan, et al. as they also noticed an increase in corrosion in Ti alloy produced using SLM and post-build heat treated to 900 °C and 1100 °C. However the corrosion was not a function of temperature here as the specimen treated at 900 °C produced more corrosion than the 1100 °C and non-treated sample exhibited better corrosion resistance (Chandramohan et al. 2017). Chen et al. compared the corrosion properties of SLM Ti6Al4V and wrought Ti6Al4V in Hank’s solution and found that additive-manufactured product is better than their counterpart (Chen et al. 2017). The use of simulated body fluids (SBF) is expected to produce results that resemble close to in vivo conditions. Xu et al. studied corrosion as a function of heat treatment and laser energy density on SLM Ti6Al4V with SBF as the electrolyte. The electrochemical corrosion test revealed that the heat-treated workpiece and the annealed workpiece is showing a low tendency for corrosion than the wrought sample which is contradictory to previous test results (Xu et al. 2017). Chiu et al. used the electrochemical method with SBF as the electrolyte to analyse the influence of process parameters on corrosion and tribo corrosion in SLM and wrought Ti6Al4V. They used different laser scan speeds, hatch distances and energy densities to obtain different surface energy densities. They did not observe any noticeable difference in corrosion rate among samples, but the tribo corrosion behaviour was different. The corrosion-to-wear ratio (C/W ratio) of wrought Ti6Al4V was lower than SLM Ti6Al4V since the wear on the wrought sample was very high (Chiu et al. 2018). Zhao et al. compared the corrosion behaviour of SLM and EBM Ti6Al4V with wrought one. It was proven that the SLM product was superior in overall corrosion resistance (Zhao et al. 2017). Avi Leon et al. studied the effect of strain rate on stress corrosion in SLM Ti6Al4V and compared it with a wrought alloy of grade 5. They used open circuit potential (OCP), potentiodynamic polarization (PD) analysis and impedance spectroscopy (EIS) to study normal corrosion and stress corrosion behaviour was examined at various strain rates using slow strain rate testing (SSRT) in a 3.5 % NaCl solution at ambient temperature. They confirmed that the SLM product was more corrosion resistant and stated that the porosity of the printed product due to inherent imperfections could have influenced the result. But this imperfection had a negative influence on stress corrosion as wrought products produced more ductility and took longer to fail (Leon et al. 2020). This could be due to the different microstructure and thermal histories of additive manufacturing that lead to defects like surface roughness or pores defect concentration or residual stresses (Kong et al. 2019). SLM Ti6Al4V corrosion resistance is being tested in salty conditions to review its application in seawater. The results proved the excellent corrosion resistance of Ti alloy once again even though additive manufactured products have some inherent defects (Bower et al. 2020).

4 Tribocorrosion

Tribocorrosion, also known as wear-accelerated corrosion or corrosion-accelerated wear, is the combined effect of wear and corrosion on the implant. Here, the wear increases the corrosion by reducing the corrosion resistance of the material or corrosion paves the way for increased wear by making more contact area or increasing the friction due to reduced surface smoothness. Tribocorrosion is a constant change due to chemical, mechanical, or electrochemical properties under a variety of conditions (Landolt et al. 2001; Mischler et al. 1999). A sliding contact pair is a major source of tribocorrosion. A stable passive film, like an active oxide film, can protect the material from tribocorrosion up to an extent. But the mechanical movement can completely or partially destroy this film (Rossi et al. 2000). Common biomaterials showed changes in current when electrochemical properties were used as the basis for the study, indicating that friction can influence tribocorrosion (Okazaki 2002). The synergism between wear and corrosion is very important in understanding the tribocorrosion mechanism. Ferreira et al. conducted experiments for this purpose on titanium alloy in Ringer solution and derived some valuable conclusions. They found that higher sliding contact speeds will facilitate more and more mechanical wear loss in the material and more wear always invites associated corrosion. The synergistic factor increases with an increase in load and speed. The wear always accelerates the corrosion at any load or speed condition whereas the corrosion inhibits wear at all conditions except high speed and high load at which the synergism is the highest (Ferreira et al. 2018). The importance of studying tribocorrosion is evident from the investigation conducted by Mathew et al. and it is illustrated in Figure 1 He established that the main reason behind implant failures is the synergistic effect between corrosion and wear than their individual effects (Mathew and Wimmer 2011).

Figure 1: 
					Synergistic component of the total weight loss under potentiostatic tribocorrosion test. Reproduced with permission from (Mathew and Wimmer 2011). Copyright 2013 Elsevier Books.
Figure 1:

Synergistic component of the total weight loss under potentiostatic tribocorrosion test. Reproduced with permission from (Mathew and Wimmer 2011). Copyright 2013 Elsevier Books.

The tribocorrosion analysis on commercial pure titanium (CP–Ti), Ti6Al4V alloy, pure niobium (Nb), and a Co28Cr6Mo alloy using a Zirconia ball under artificial saliva condition demonstrated that the Co28Cr6Mo alloy has the better tribocorrosion resistance and wear resistance while the CP Ti had superior corrosion resistance (Vilhena et al. 2019). The reason for the deviation in corrosion behaviour is due to the formation of highly stable oxide films on Ti and Ti alloys. Ponthiaux et al. concluded that the wear can cause a galvanic couple between worn and unworn areas (Ponthiaux et al. 2004).

Steels, cobalt, and titanium-based alloys are the most used metallic biomaterials (Elias et al. 2008; Li et al. 2015; Yazdi et al. 2018). Recently, titanium-based alloys are widely used compared to others due to their comparatively superior desirable properties. It has good corrosion resistance and high strength even though it is a lightweight material. Physical, mechanical and corrosion resistances are also in favour of titanium (Bailey 2018; Gao et al. 2018; Huang et al. 2015; Yazdi et al. 2017). But they do not have good resistance towards wear, and tribological and fretting responses are also poor. Researchers found that this could be due to the low thermal conductivity of titanium (Chauhan and Dass 2013; Licausi et al. 2015; Pejaković et al. 2018). This could lead to the continuous removal of the oxide layer which accumulates metal ions and wear debris in the blood and cause biological reactions which in turn could fail implants (Dimah et al. 2012).

A comparison of tribocorrosion of traditionally forged and SLM-printed Ti6Al4V was made by Huang et al. who found that the forged sample has better corrosion resistance while the SLM sample fares better in wear tests. Hence, the performance of the SLM sample should be superior under normal conditions; however, if the point of application is corrosive in nature, such as bone scaffolds, the forged samples would perform better than the counterpart (Huang et al. 2022). The uniform structure and high β phase content in the forged samples were explained as the reason for the better corrosion resistance. The role of structure in tribocorrosion was studied using AlSl 316L stainless steel and results were corroborative with the previous conclusion as the formation of gradient nano-structured surface felicitated the formation of pits even before the passivation voltage (Zou et al. 2022).

There are many challenges faced by scholars as it is a relatively new field of research. There are fewer references is one of them and the lack of a recognized standard test apparatus is another (Mathew et al. 2009). The electrochemical properties of the surface before and during the friction process are analysed to quantify the synergism between wear and corrosion (López-Ortega et al. 2018). ASTM G99 and ASTM G133 are 2 universally accepted test protocols for wear test, which uses pin-on-disc unidirectional circular configuration ball on flat unidirectional linear configuration, respectively (Azzi and Klemberg-Sapieha 2011). The standard corrosion measuring techniques such as potentiodynamic, potentiostatic and electrochemical impedance spectroscopy (EIS) are combined with tribological tests to evaluate the tribocorrosion behaviour. An ASTM standard for tribocorrosion was specifically introduced in 2004, combining the tribological and electrochemical loading in the same apparatus, but was later retracted in 2016 amid criticism from several authors. However, the American Society for Testing and Materials (ASTM) reapproved this standard guide to determine synergism between wear and corrosion in 2021 which could bring uniformity to the testing procedure (ASTM G119-09, 2021). Two approaches are in practice to calculate the material deterioration where the wear is carried out in controlled potential using external current it is carried out at open circuit potential (OCP).

Figure 2 represents a standard tribocorrosion measurement setup. It can simultaneously measure wear as well as corrosion using electrochemical parameters (Anaee and Abdulmajeed 2016). It consists of 3 electrodes, namely the working electrode, reference electrode and auxiliary electrode. The working electrode is the sample on which we are performing the tribocorrosion test. The reference electrode is the electrode with a known potential used to compare the potential of the working electrode. Standard Hydrogen Electrode (SHE) is considered the best reference electrode with a potential of 0 V. However, standard calomel electrode (SCE) or silver chloride electrode (AgCl) saturated with KCl or NaCl is used due to the difficulty in handling hydrogen. The auxiliary electrode, also known as the counter electrode is used to allow current to flow in the electrochemical cell and reduce the current flow in the reference electrode. Platinum is the best counter electrode available due to its inertness. However, carbon, copper or stainless steel can also be used depending on the application.

Figure 2: 
					Schematic representation of the tribometer used to conduct tribocorrosion tests. Reproduced with permission from (Huang et al. 2022) Copyright 2022 Elsevier.
Figure 2:

Schematic representation of the tribometer used to conduct tribocorrosion tests. Reproduced with permission from (Huang et al. 2022) Copyright 2022 Elsevier.

Many researchers made attempts to improve the tribocorrosion resistance of biomaterials. They found that surface treatment can improve the property considerably. They also tried coating the substrate (biomaterial) with different components. Some of them were successful and proved to be very effective. These coatings have the advantage of providing osseointegration over surface treatment. Nevertheless, different approaches are still being made in the attempts to improve tribocorrosion resistance. The most significant and recent techniques for tribocorrosion resistance improvement are discussed in the following sections.

5 Controlling techniques for tribocorrosion

Since tribocorrosion is the combined effect of corrosion and wear, we should individually control corrosion or wear or both to reduce the impact of tribocorrosion in implants. There are some common techniques adopted to control corrosion in metals which can resist the inevitable to an extent. These general engineering methods are not always applicable to biomedical applications due to toxicity and biocompatibility issues and care must be taken in selecting the technique to ensure that it does not cause any threat to the host.

Surface modifications are generally adopted in resisting corrosion and wear in biomaterials. It is classified into 2 major categories, (i) physicochemical modifications which alter the atomic or molecular structure or modify the compound structure on the surface and (ii) surface coatings which use different materials that adhere to the material surface and protect from surrounding reactions. Physicochemical modifications include chemical reactions, etching, and mechanical roughening/polishing and patterning whereas the coating technique comprises grafting, non-covalent and covalent coatings, and thin film deposition. Physicochemical modification changes are applied on the surface with a thickness level range of 10 nm–100 nm which ensures the mechanical and chemical stability of the material (García 2011).

Table 1 shows the different methods adopted and their outcomes on the most popular implant materials, Titanium and Titanium alloys. Other popular methods used on Ti and other materials are explained in the following sections.

Table 1:

Different modification methods attempted on Ti and its alloys to control tribocorrosion.

Sl. no. Substrate material Modification material Methods Results References
1 Ti Zr, Hf, Ta, Nb Alloying Corrosion and tribocorrosion resistance improved. Reduced defects in passive films (Wang et al. 2022)
2 Ti, Ti6Al4V Nb, Al Powder metallurgy The fabrication process of titanium alloys influences the corrosion and tribocorrosion behaviour of those biomedical alloys (Licausi 2017)
3 Ti6Al4V Alumina, Yttria stabilized Zirconia, Titania Plasma spray coating Prevented tribocorrosion and only Wear loss were present (Cheng et al. 2019)
4 Ti, Ti6Al4V and Ti13Nb13Zr Nanocrystalline diamond (NCD) Chemical vapour deposition (CVD) Even though corrosion was reduced, Tribocorrosion performance was poor due to the penetration of electrolytes through the pores (Gopal et al. 2017)
5 Ti6Al4V Poly-ether-ether-ketone (PEEK) Hot isostatic pressing (HIP) PEEK protected the surface against tribocorrosion and produced a lubricating effect which resulted in a specific wear rate much lower than the bare sample (Sampaio et al. 2016)
6 Ti6Al4V PEEK Hot pressing Improved tribocorrosion behaviour while maintaining the load-bearing properties (Bartolomeu et al. 2019)
7 Ti6Al4V Hydroxyapatite (HAP) Hot pressing The COF was increased with an increase in HAP content, but the specific wear rate, corrosion and tribocorrosion were reduced by reducing the abrasive and adhesive actions (Buciumeanu et al. 2017)
8 Ti6Al4V Diamond-like carbon (DLC) CVD The coating is effective against corrosion and wear individually in static conditions, but not efficient in tribocorrosion conditions (Manhabosco and Müller 2009)
9 Ti Copper (Cu) Alloying Ti Cu alloy performed better in tribocorrosion performance due to its better repassivation property and hardness (Bao et al. 2018)
10 Ti Various porosity levels Powder metallurgy Porous samples were unable to repassivate like the dense sample and hence produced detrimental effects on material degradation (Manoj et al. 2019)
11 Ti6Al4V Poly (D,L-lactic acid) (PDLLA) Spin coating The coating provided protection against tribocorrosion, but the durability of protection is very small (Souza et al. 2015)
12 Ti6Al4V Oxygen diffusion layers (ODL) Thermal oxidation Even though the corrosion behaviour was similar, the hardness of ODL provided better protection against wear and tribocorrosion (Yazdi et al. 2017)

5.1 Coatings

This is the most common corrosion control technique. The coating layer will isolate the metal structure from the corrosive environment (Yan 2006). Wear is caused by the mechanical contact between 2 surfaces, and it progresses gradually. Tribology is a branch of science that deals with the study of wear as well as friction and lubrication. Lubrication is provided to reduce the rate of progress of wear. Lubrication will separate the two mating surfaces with a thin layer of film. The applied load is carried by pressure generated within the fluid, and frictional resistance to motion arises entirely from the shearing of the viscous fluid (Stachowiak and Batchelor 2001).

5.1.1 Bioceramic coatings on implants

These types of coatings were primarily introduced to improve the osseointegration and related biological properties which later proved to be efficient in controlling corrosion too. They control the problems due to the difference in mechanical properties of metals and natural bone, prevent the release of metal ions to the body to an extent and facilitate the regeneration of natural bone. Hydroxyapatite (HAP), tricalcium phosphate (TCP) etc. are the commonly used ceramic coatings in orthopaedics (Balamurugan et al. 2008). Kaur et al. found that titanium improved corrosion resistance by 73 % and stainless steel by 51 % when they were coated with hydroxyapatite (HAP) using the biomimetics method. The samples maintained the surface morphology intact too (Kaur et al. 2019). Czechowska et al. pointed out that the use of HAP with chitosan (CTS) increased the adhesion or bond strength of the HAP layer over the substrate (Czechowska et al. 2016). Gayathri et al. found that the corrosion resistance of Ti was increased considerably after coating with magnesium incorporated HAP using the hydrothermal method in Ringer’s solution and its performance was better than Ti with only HAP coating (Gayathri and Bhupathi 2019). Tohidi et al. developed a HAP coating on NiTi alloy using pulsed electrodeposition in an optimized magnetic field. They found that the HAP-coated NiTi produced better corrosion resistance compared to uncoated NiTi and heat-treated NiTi (Tohidi et al. 2020). They concluded that this improvement was due to the reduction in localized and pitting corrosion and coating done at optimized conditions facilitates uniform corrosion which avoids sudden failures of the implant. D. Gopi et al. conducted corrosion and mechanical properties analysis on a high energy low current DC electron beam (HELCDEB)-treated titanium (Ti) coated with mineral substituted (strontium, magnesium and zinc) HAP (M-HAP). Test results revealed that M-HAP coated samples made a more positive shift towards nobleness. HELCDEB sample also produced improvement in corrosion resistance over the untreated sample. The energy density of the EB machine also influenced the final properties as the 700 KeV energy density treated sample was better than the 500 KeV treated ones (Gopi et al. 2014). Priyadarshini et al. analysed the effect of multi ionic doped porous HAP (Ce4+/Si4+ doped HAP) coating on corrosion resistance and biocompatibility of Ti6Al4V. The coating produced using the sol-gel method at 4000 RPM on a heat-treated substrate for 2 h at 500 °C produced significant improvement in corrosion resistance and biocompatibility (Priyadarshini and Vijayalakshmi 2021). The Analysis of the influence of HAP coatings on corrosive tendency developed using the pulsed laser deposition method on SS316L and Ti6Al4V revealed the following facts along with biocompatibility. The test was carried out in Hanks’ solution and the data collected revealed that HAP-coated Ti6Al4V shows better corrosive characteristics than SS316L (Ponnusamy and Muthamizhchelvan 2018). The nanocomposite coating of HAP + TiO2 using electrophoretic deposition on Ti13Nb13Zr was proven to be a good technique to improve the corrosion resistance of the implant material which is similar to the result obtained in plasma spraying technique coatings (Koike and Fujii 2001; Mohan et al. 2012). The electrophoretic deposition reduced the corrosion by around 20 times lower than the uncoated sample as the corrosion reduced from 0.02692 may to 0.00145 mpy. Zn-doped HAP coatings also produced excellent corrosion resistance. But the authors did not give qualitative data regarding the improvement made from the original sample (Take et al. 2014). A novel plasma spraying method was developed by Hameed et al. and named as axial suspension plasma spraying (SPS) to coat HAP onto Ti6Al4V in order to overcome some of the demerits of commercially used and only recognized atmospheric plasma spraying (APS) technique such as loss of crystallinity and poor adhesion strength. Instead of propelling molten HAP particles as in APS, a suspension was made with HAP in either water or ethanol and fed axially to the plasma plume. The use of suspension reduced the operating temperature since HAP need not be melted which aids in maintaining the crystallinity of the HAP. The suspension feed is also useful in producing a coating thickness of around 50 μm while the APS technique produces coatings with a thickness of around 100 μm. While they were successful in achieving their objective, they also tested the sample for corrosion resistance in 10 % fetal bovine serum (FBS). The corrosion potential (Ecorr) and corrosion current density (Icorr) obtained from the potentiodynamic polarization test showed that the sample coated with the SPS technique displayed the highest negative Ecorr and least Icorr values which proves that the corrosion resistance is better than the contemporary technique and authors claimed that this is 9.5 times higher than the currently used APS sample (Hameed et al. 2019). A combination coating of HAP and chitosan (CTS) was used on SS 316L to analyse the tribological properties. The electrodeposition method was adopted for coating and 3 different coating periods (15, 22.5 and 30 min) were tested in the experiment. The results obtained derived a conclusion that the coating improved the coefficient of friction and frictional force of the sample and it helped to reduce the wear rate of the sample. The longer time period of a deposition provided more suitable outcomes as they did not get detached from the substrate surface at lower load tests (García et al. 2020). This coating along with PEEK has already proven to be biocompatible with better anti-bacterial properties (Abdulkareem et al. 2019). Even though ceramics are corrosion-resistant by nature, their low impact resistance, brittleness, difficulty, and cost of production reduce their wide usage in orthopaedics (Sin 2015). However, they are used to coat the parent metal to induce their desirable properties in the implant material (Metikoš-Huković et al. 2006; Stack et al. 2010). Cheng et al. studied the effect of plasma-sprayed hard ceramic coatings on the tribocorrosion behaviour of Ti6Al4V alloy. Monolithic micron alumina (IDA), micron alumina −40 wt % yttria-stabilized zirconia (YSZ) composite coating (IDAZ) and by-layer nanostructured alumina −13 wt % titania/YSZ (IDZAT) were coated on Ti–6Al–4V alloy and the thickness of coatings was approximately 250 µm for IDA and 400 µm for IDAZ and IDZAT. The samples with coatings did not experience any tribocorrosion effect and only pure wear was present. The wear loss in comparison with the bare substrate was too small to measure. The IDZAT produced the best performance by exhibiting more positive potential, higher impedance value and lower COF value as illustrated in Table 1 (Cheng et al. 2019). Poly-ether-ether-ketone (PEEK) was veneered to Ti6Al4V and the effect on tribocorrosion was analysed using alumina balls. The COF dropped to 0.07 on coated sample from 0.36 on the uncoated sample and hence the tribocorrosion performance was also improved which is mentioned in Table 1 (Sampaio et al. 2016).

5.1.2 Diamond-like carbon coatings (DLC)

Diamond-like carbon (DLC) is a chemically inert, extremely hard, wear-resistant and biocompatible material used to coat implants. They reduce the degradation rate of polythene cups used in hip joints by a minimum factor of 10 and by a factor of 105 in sliding contact metal joints with a coating of thickness 1 µm (Tiainen 2001). DLC coatings improve the corrosion resistance of metallic biomaterials by maintaining their biocompatibility, inducing high hardness, low coefficient of friction, making the surface chemically inactive and increasing electrical resistivity and most importantly, high wear resistance (Love et al. 2013). Ti6Al4V coated with DLC using CVD technique provided excellent wear resistance but did not last long as mentioned in Table 1. This could be due to the loss of adhesion between the parent metal which is caused by the porous nature of DLC (Manhabosco and Müller 2009). The adhesion problem issue was further noticed by Azzi et al. when they found that DLC coatings had reduced the adhesion performance of stainless steel compared to other coatings (Azzi et al. 2009, 2010). The DLC coating works better with titanium-based alloys than stainless steel and CoCrMo substrates (Zhang et al. 2015). The corrosion and tribocorrosion behaviours of titanium alloys (60NiTi and Ti6Al4V) were examined in synthetic urine conditions with and without DLC coating with an average thickness of 2.19 µm. All coated samples have displayed a reduced and stable coefficient of friction, unlike the bare sample. The heat-treated 60NiTi (NiTi60T) model produced better tribocorrosive properties than the other samples OF 60NiTi and Ti6Al4V. The DLC coating reduced the corrosive nature by 70 times on 60NiTi and similar effect on other samples too even if the magnitude of impact is low (Paula et al. 2018). Similar effects were noticed by Zhao et al. when they tested the effect of DLC coating on SS316, and CP Ti for tribocorrosion and they were unable to detect any wear loss on the samples in their 90-min test. However, the coating was not that effective in CoCrMo alloy (Zhao et al. 2016a).

A new technique known as microarc oxidation (MAO) or the plasma electrolytic oxidation (PEO) process is carried out on the substrate before coating with DLC to eliminate the mismatch between coating and sample. MAO is used to increase the wear resistance, corrosion resistance and load-carrying capacity of titanium and its alloys. This duplex coating provided on Ti6Al4V produced the highest corrosion potential and lowest corrosion current than the individual coatings along with a stable and low coefficient of friction and desirable morphological structure (Sukuroglu et al. 2015).

5.1.3 Other coatings

Tao Shao et al. analysed the influence of Si content on the tribocorrosion behaviour of Cr1-xSixN coatings prepared using the magnetron sputtering technique. They found that with the increase in Si content, the corrosion resistance increases, and the wear resistance first increases and then decreases. The coating with a combination of Cr0.58Si0.42N exhibited the best corrosion resistance and wear resistance. However, the tribocorrosion performance was not significantly improved at this combination as expected and the authors claimed the reason for this phenomenon was embrittlement resistance in salt water environment (Shao et al. 2018). Manjunath et al. conducted studies on the effect of Ti and tetrahedral amorphous carbon multi-layered alternate coatings up to a thickness of 936 nm on silicon and stainless steel 202 on corrosion and wear. The reciprocating wear studies recorded a reduction in the coefficient of friction ranging from 1/10th to 1/4th at different load conditions in coated SS compared to the bare substrate. It also enhanced the corrosion resistance since the corrosion rate reduced from 0.017 Mils penetration per year (mpy) to 0.007 mpy (Manjunath et al. 2017). Tribocorrosion characterisation of the NiTiCu alloy, a copper blended NiTi alloy, produced a reduction in coefficient of friction (COF) by 20 % approximately. They also found that the presence of 5 % copper produced better tribocorrosion results than the 10 % copper sample since the wear loss in the latter sample was more than in the former whereas friction properties were almost similar (Sharma et al. 2019). But it is unable to confirm whether 5 % is the optimum ratio of carbon for better performance since they did not test any other proportions of alloys in testing. pulsed-DC plasma-assisted chemical vapour deposition (PACVD) method was used to deposit TiSiN nanocomposite films on Ti6Al4V by Movassagh–Alanagh et al. to study the influence of Si content on wettability and corrosion resistance. They found that the corrosion resistance improved considerably and concluded that this is due to reduced microstructural defects that occurred as a result of the addition of Si to TiN which forms nanocomposite-type microstructure. This layer also houses a hybrid TiO2/SiO2 passive oxide layer (Movassagh-alanagh et al. 2018). Trino et al. manufactured the body and surface of the Ti biomaterial separately where the Ti surface was modified with TiO2, two different spacers, 3-(4-aminophenyl) propionic acid (APPA) or 3-mercaptopropionic acid (MPA) and dentin matrix protein 1 (DMP1) peptides. All coated samples presented improved corrosion resistance and TiO2 was the best among all due to the presence of a protective oxide layer. However, the tribocorrosion performance of TiO2 APPA P and TiO2 MPA P, significantly improved which could be due to the ability of peptides in mimicking the natural tribolayer, which consists of tribochemical products and a carbonaceous material capable of reducing the mass loss. This indicates that the presence of DMP1 enhances the utility of Ti implants (Mathew et al. 2014; Trino et al. 2018). Liu et al. investigated whether graphene coating using a modified chemical vapour deposition (CVD) could prevent the release of carcinogenic Ni2+ ions to the body from the popular shape memory biomaterial NiTi. They found that the dangerous ion release reduced from 134 μg/L to 46 μg/L from the pristine sample to coated sample in the first 24 h and the trend continued for long-term observation too. The coated sample also exhibited better corrosion resistance as corrosion potential shifted positively to −0.18 V from – 0.43 V with standard calomel electrode (SCE) as the reference (Zhang et al. 2016). An attempt to improve the wear and corrosion behaviour of Ti by developing TiC structure as a coating on the surface using laser surface treatment by Mohazzab et al. Wear test, impedance spectroscopy and polarization analyses revealed that the process made positive impacts and made material better than pure Ti with optimum ablation time. Prolonged ablation reduced coefficient friction as a function of time, but wearing tendency inversed after some time which could be due to unevenness of the TiC layer formed due to overexposure to heat. Even though the laser-treated samples displayed a positive shift in electrode potential, the corrosion performance was almost similar to pure Ti (Mohazzab et al. 2020).

In an experiment to identify if Ti15Mo alloy could be used as dental implant biomaterial, the effect of pH and fluoride content, which usually varies in the dental vicinity due to food intake, on the corrosiveness of the material was analysed. This is relevant to orthopaedics application also since the pH in the implant vicinity could change due to foreign body reactions of the human body’s immune system. The tests were carried out at variable pH values ranging from 3.5 to 7.2 and fluoride ion concentrations from 2500 to 10000 ppm. The increase in fluoride content resulted in increased corrosion current (Icorr) indicating the negative influence. The acidic nature (pH 3.5) also produced damaged surfaces suggesting a negative influence on material stability. However, The Ti15Mo survived the minimum requirements in all test conditions and could be used for producing dental implants (Sasikumar and Rajendran 2018).

It is well known that alloys with passivation (formation of the metal oxide layer of a few nanometer thicknesses) have better corrosion resistance than active metals. Bidhendi et al. compared the corrosion performance of commercially pure titanium, (CP Ti), TI6Al4V and SS316LVM in Hank’s solution. They found out that CP Ti has the lowest corrosion rate while SS 316LVM performed the worst coupled with localized corrosion in bare samples without any treatment. Passivating the SS sample has somehow reduced the corrosion issue. Anodizing the Ti and its alloy produced considerable improvement in corrosion resistance as a double layer of metal oxide is formed (a porous layer on top of a thin layer). Both samples exhibited similar resistance with CP Ti having a slight edge over the other alloy sample (Bidhendi and Pouranvari 2011). Imparting molybdenum (Mo) to the shape memory alloy NiTi by laser energy is done primarily to reduce the release of toxic Ni ions to the blood which constitutes 55 % of the total weight of the alloy since it is carcinogenic, allergic and toxic in nature. Molybdenum could improve the toxicity and biocompatibility of NiTi. The process not only reduced the release of Ni into the blood, but also improved the corrosion resistance and hardness and even facilitated HAP growth indicating better bioactivity (Ng and Man 2012). TiO2 gives the best result among all coatings tested so far. A comparison with Cr2O3 coatings in saline conditions proves that TiO2 yields far better results than the former one as the former coating was found to be phased out on a large area after a 10-month inspection in ball valves (Kim and Walker 2007). The effect of load and particle concentration on tribocorrosion was studied by Stack et al. and the role of corrosion in wear is not significant at higher loads and high particle concentration as in low load and concentration. But there is no consistency in wear rate and wear is more at low loads at lower particle concentration and maximum at higher loads with high particle concentration. The authors attributed this phenomenon to a change in the wear mechanism at higher loads due to the tribo-chemical mechanism (Stack et al. 2011).

Temporary anchorage devices (TAD) are used in orthodontic treatments to fix the implants to the bone. Ti6Al4V and stainless steel are commonly used for this purpose. A comparison in corrosion behaviour of these materials in the more corrosive environment (due to variable temperature, pH etc. of consuming food) of saliva is tested to help material selection. Even though stainless steel produced a more quality surface finish, which can eliminate entrapment of impurities, corrosion and ion release were lower in titanium alloy. The reason could be the acidic treatment of the surface from the surroundings or self-passivation of the alloy and the output is in agreement with previous attempts by Suzuki et al. and Huang et al. (Huang et al. 2003; Nascimento et al. 2020; Suzuki et al. 2018). The effect of fluoride concentration (as they are present in abundance in mouth cleansing products) on Ti alloy (Ti15Mo) corrosion was studied by Sasikumar et al. at various pH values in artificial saliva to mimic the original conditions. The cathodic shift in corrosion potential from −0.107 to −0.398 against the saturated Ag/AgCl electrode at constant pH indicates the influence of fluorides in accelerating the corrosion and reduction in pH value also facilitates corrosion (Sasikumar and Rajendran 2018).

5.2 Surface treatments

Surface treatment methods like shot peening or nitriding, ion implantation, passivation etc. could reduce the corrosion rate considerably which has been proven through experiments (Buchanan et al. 1987; Maurer et al. 1993). Treatment of metals in acidic or alkaline medium, heat treatment etc. are the commonly used methods to improve the passivation of a material which is the formation of a relatively inert oxide layer on the metal surface (Deakin et al. 2004; Tamilselvi et al. 2010).

Carburisation is a process of heat treating a metal or alloy in the presence of a carbon source so that carbon gets absorbed and deposited in carbide forms on the metal surface. The corrosion and tribocorrosion behaviour of carburized commercially pure titanium (CP Ti) at 925 °C revealed that the titanium carbide layer is formed at the top surface with an oxygen diffusion zone (Bailey and Sun 2018). The corrosive property study of TiC earlier revealed that the positive potential shift in corrosion potential is not reflected in actual practice and the material loss was more than the pure sample (Oláh et al. 2015). But the novel technique employed here to reduce oxygen diffusion created multiple layer coatings with TiC, TiO2, TiO, TiCxO1−x and α-Ti. This resulted in offering great resistance to corrosion initially, but the resistance was reduced once the first layer was broken (Bailey and Sun 2018).

5.2.1 Ion implantation

Ion implantation is a traditional technique used to improve the surface hardness of a material and thereby increasing the corrosion as well as wear resistance. Buchanan et al. were the first team to introduce ion implantation for improving biomaterial life for the first time. They introduced nitrogen ions onto the surface of Ti6Al4V without affecting the bulk properties of the alloy and found that it reduced tribocorrosion compared to the normal sample (Buchanan et al. 1987). This was confirmed by Sundararajan et al. on CP Ti (commercially pure Titanium) who found that a dose between 4 × 1016 and 7 × 1016 ions/cm2 is optimum for orthopaedic applications and overdosing will produce detrimental effects which could be due to the formation of oxynitrides during the implantation (Sundararajan et al. 1999).

5.2.2 Nitriding

Nitriding is the process of diffusing nitrogen ions through heat treatment into the substrate surface so that it can form nitrides which reduce the reactivity of the same. It will also improve the hardness of the surface (Guo et al. 2015; Zhecheva et al. 2005). The thickness of the protective layer can be increased by increasing the nitriding temperature, which in turn improves the wear resistance (Farè et al. 2012). Azzi and Szpunar found that nitriding the pure titanium at 700 degrees Celsius made the material nobler as it bettered the resistance towards corrosion in the ringer’s solution. It is a well-known fact that titanium alloys are protected by titanium oxide layers on the nanometre level (Azzi and Szpunar 2007). Tribocorrosion and crevice corrosion can be reduced by using materials with low surface roughness (Hammood et al. 2019). The presence of the TiN layer improved the tribocorrosion resistance compared to the bare metal (Azzi and Szpunar 2007). Nitriding at low temperatures on SS316 may not be sufficient to improve tribocorrosion resistance even though it improves wear resistance. High-temperature nitriding also produced a similar effect as the samples became more sensitive towards anodic corrosion in PBS solution. On the other hand, nitriding produced positive results on CP Ti samples and their corrosive, wear and tribocorrosion resistance was improved considerably (Zhao et al. 2016a). Hence the corrosion rate can be controlled. Venkatesan et al. used biomaterials SS316L and Ti6Al4V for plasma nitriding for 8 h at 480 °C and 24 h at 520 °C on different samples and tested the corrosion performance. 3.5 % NaCl solution was used as the electrolyte with reference electrode AgCl/Ag and Counter electrode platinum wire. They concluded that the 24 h heat treatment improved the corrosion resistance of both materials more than the 8 h heat treated one. The corrosion resistance of SS316L was better in 24 h samples whereas Ti6Al4V displayed better performance in 8 h samples (Venkatesan et al. 2021). They did not mention whether it is the duration of treatment or the temperature that influenced the corrosion performance.

One of the major drawbacks of TiN coating is that it has an oxidation tendency. The addition of Al to TiN (TiAlN) overcomes this issue by forming Al2O3 layers at high temperatures. The PVD technique was used to produce a coating of TiAlN on stainless steel (AISI410) and potentiodynamic polarization methods were used to study the corrosive behaviour relative to the bare substrate. The test was conducted in 3.5 % NaCl solution and the coated sample produced a 34.51 % protection efficiency which is more than 3 times the uncoated sample and the corrosion rate was reduced to 0.3923 mpy from 1.7956 mpy (Cunha et al. 1998; Prabakaran 2019).

5.2.3 Passivation

Passivation is a technique used to protect metals and alloys from corrosion where an outer layer to a material to protect it from harmful reactions. This passivation can be either self-passivating where a layer of metal oxide is formed at the outer surface of the substrate, just like the formation of TiO2 on the surface in the case of titanium alloys or applying a micro coating on the material surface to isolate it from the environment.

The surface passivation of SS316 L with nitric acid enriched with Cu ions not only improved the corrosion resistance of the material but the antibacterial performance was also improved by the deposition of Cu along with improved biocompatibility (Zhao et al. 2016b). The chemical passivation of Co Cr-based dental alloy also produced a reduction in pitting corrosion, but the pH value of the passivating medium does influence the change in properties with low values improving the corrosion resistance (Rylska et al. 2017). Titanium and its alloys are highly reactive with oxygen that a new layer will be spontaneously formed whenever the existing layer is deteriorated due to mechanical wear eliminating the need for manual treatment. Other than forming this protective layer, passivation helps maintain the surface roughness and the homogeneity of the material by removing other inclusion in the material.

5.2.4 Laser surface treatment

Surface treatment using an excimer laser by Yue et al. reduced the corrosion rate by 7 times compared to the untreated sample of Ti6Al4V (Yue et al. 2002). Laser surface treatment has many advantages over other techniques such as high speed, high adhesion strength, complex shape coating etc. different laser sources are used and CO2, YAG, excimer, dye, argon-ion, diode etc. are commonly used depending on the requirement of the coat (Richard et al. 2010).

5.2.5 Electropolishing and thermal methods

Many researchers were successful in improving the corrosive behaviour of biomaterials using polishing and thermal treatments which are cost-effective compared to other methods.

5.3 Electrochemical methods for corrosion resistance

Here the implant material is immersed in an electrolyte and is connected to an electrical circuit. Electrochemical methods are categorized into 3, namely anodization, electropolishing and electroerosion (Richard et al. 2010). Anodization, otherwise known as anodic oxidation is used to produce improved oxide layers or improved oxide layer thickness up to micron level and helps in producing a surface with complex topography, and porosity which helps in corrosion protection and reduced release of ions into the body.

5.3.1 Cathodic protection (CP)

As discussed in the previous sections, corrosion can be treated as an electrochemical cell reaction. In this method, the metal surface to be protected is made the cathode of an electrochemical cell to avoid anodic dissolution.

5.3.2 Anodic protection (AP)

If the metal to be protected cannot be made the cathode, provides some protective or passive film with the help of externally applied anodic currents.

5.4 Biodegradable/absorbable metals

Materials which degrade gradually in vivo conditions, but do not cause any long-term effect on the body (ASTM F3160-16). Even though these materials do not reduce corrosion or wear, the by-products are biocompatible and not toxic or dangerous to the host body in any manner. Iron (Fe), magnesium (Mg) and zinc (Zn) are the commonly used absorbable metals (Li et al. 2020; Wang et al. 2020) and zinc is considered as the better choice due to its electrochemical properties. Its standard electrode potential against standard hydrogen electrode is having an optimum value of (−0.76 VNHE) which is neither high as Mg (−2.37 VNHE) nor low as Fe (−0.44 VNHE) indicating optimum corrosion potential (Chen et al. 2016; Escobar et al. 2019; Mo et al. 2021; Wang et al. 2019). These types of materials are not directly used as implants but as the support needed in cases like orthopaedic and cardiovascular applications. They should provide support for a minimum period of 3 months and a maximum of 2 years within which it should be dissolved completely. The duration of support and hence the ideal corrosion rate depends on the location of the implant (DebRoy et al. 2018; Francis et al. 2015; Venezuela and Dargusch 2019; Zheng et al. 2014). The corrosion products of biodegradable should also be biocompatible along with the parent metal. All these 3 biodegradable metals are present in the human body indicating their biocompatibility. Zn and Mg are stored in the inorganic bone matrix whereas Fe is used to produce haemoglobin. However, a healthy diet demands more Mg intake than Zn or Fe. Hence it is better to use Mg as a suitable material for biodegradable implants, but it is inhibited by its corrosion potential. Additive manufactured Mg implant lost 20 % weight in 28 days in in-vivo conditions while the Fe lost 3 % and Zn lost 8 % only (Habibovic and Barralet 2011; Li et al. 2018; Marie 1992). Efforts have been made to improve the corrosion resistance of Mg. The addition of 1 % Tin (Sn) improved corrosion performance while its increased percentage reduced the resistance. The authors concluded that the reason for this could be grain refinement is not sufficient to overcome the additional phases produced by the extra addition of Sn (Zhou et al. 2016). Similar results were obtained on another version of Mg alloy (ZK30 with a combination of Mg–3Zn-0.6Zr) (Shuai et al. 2018a). Alloying Mg with Al was proven to be an effective way to control corrosion, but reduced biocompatibility which prevents its use as an implant (Ghosh et al. 2020). Zn is the metal with the optimum corrosion potential than Mg and Fe. When silver was added to it, the corrosion rate was increased while Mg produced both effects on the corrosion rate which was unexpected. The dependency of the concentration of Mg on grain refinement was attributed as a possible explanation for the phenomenon (Mei et al. 2020; Shuai et al. 2018b; Yang et al. 2018).

Carbon nano tube (CNT) coatings were made on the Ti surface using electrophoretic deposition to analyse the effect on mechanical and biological properties. The coating shielded the sample against corrosion and enhanced the cell viability of the material (Dlugon et al. 2015). A Ti-CNT metal matrix composite was prepared and samples were produced using powder metallurgy. The developed structures produced better resistance towards wear and corrosion than the normal sample, but the biological properties were not tested. Moreover, processing Ti and CNT at high temperatures resulted in an interfacial phenomenon which is the formation of titanium carbides (TiC) distributed all over the sample which could compromise the structural properties (Munir et al. 2015). This could be solved by adopting additive manufacturing techniques to develop the samples. The undesirable corrosive nature of SLM developed samples can be improved by annealing the samples. The SLM Ti6Al4V samples annealed at 800 °C produced comparable corrosive behaviour to a traditionally manufactured sample. This is attributed to the stress relief of the martensitic phase and the formation of β Ti6Al4V (Ettefagh et al. 2019). The corrosive, wear, tribocorrosive resistance and biocompatibility of the implant can be improved using suitable surface modification techniques such as bioceramic coating, ion implantation, electropolishing etc. (Gorejová et al. 2023; Hryniewicz et al. 2009; Kumar et al. 2019; Rautray et al. 2011; Surmeneva et al. 2020)

6 Summary

Failure of artificial implants due to corrosion and wear before the expected life period has been a concern for biomedical implants. Implants require extended service life to meet the demands of increased life expectancy of elderly patients with orthopaedic issues and young patients who are affected by accidents. A plethora of efforts were made to improve the service life of orthopaedic and dental implant materials to prevent the necessity of revision surgery. To address the above issues, strategies like tailoring of existing materials bulk characteristics, modification of existing materials surface characteristics using the latest engineering techniques etc. are adopted. The identification of tribocorrosion or synergy between wear and corrosion has helped the scientist to come up with better ideas and techniques to resolve the issue and this area is still in its infancy period which will hopefully produce tremendous outputs in the near future.


Corresponding author: Renold Elsen Selvam, School of Mechanical Engineering, Vellore Institute of Technology, Vellore, Tamil Nadu632014, India, E-mail:

About the authors

Umanath Puthillam

Umanath Puthillam is pursuing his doctoral degree in mechanical engineering at Vellore Institute of Technology (VIT), Vellore. He has a Master’s degree in production engineering and a Bachelor’s degree in automobile engineering. He has 6 years of experience in teaching at graduate level and authored 3 articles in Q1/Q2 journals.

Renold Elsen Selvam

Renold Elsen Selvam has published 41 papers in indexed journals, 4 book chapters, and edited 2 conference proceedings. He has also registered 8 patents, 2 of which have been granted. He has secured and completed multiple projects and consultancy work from national and international funding agencies, amounting to 1.5 crores.

Acknowledgments

The authors acknowledge the support and facilities provided by VIT, Vellore.

  1. Research ethics: Not applicable.

  2. Author contributions: Umanath Puthillam and Dr Renold Elsen went through multiple journals and shortlisted suitable papers. Umanath Puthillam and Dr Renold Elsen studied different journal papers and collected data. The information collected was comprehended and analysed and the manuscript was drafted by Umanath Puthillam. Dr Renold Elsen reviewed the paper.

  3. Competing interests: The authors state no conflicts of interest.

  4. Research funding: None declared.

  5. Data availability: Not applicable.

References

Abdulkareem, M.H., Abdalsalam, A.H., and Bohan, A.J. (2019). Influence of chitosan on the antibacterial activity of composite coating (PEEK/HAp) fabricated by electrophoretic deposition. Prog. Org. Coat. 130: 251–259, https://doi.org/10.1016/j.porgcoat.2019.01.050.Search in Google Scholar

Al-Mobarak, N.A., Al-Mayouf, A.M., and Al-Swayih, A.A. (2006). The effect of hydrogen peroxide on the electrochemical behavior of Ti and some of its alloys for dental applications. Mater. Chem. Phys. 99: 333–340, https://doi.org/10.1016/j.matchemphys.2005.10.032.Search in Google Scholar

Alvarez, K. and Nakajima, H. (2009). Metallic scaffolds for bone regeneration. Materials 2: 790–832, https://doi.org/10.3390/ma2030790.Search in Google Scholar

American Society for Testing and Materials (2021). Standard guide for determining synergism between wear and corrosion, ASTM G119-09 (2021).Search in Google Scholar

Anaee, R.A.M. and Abdulmajeed, M.H. (2016). Tribocorrosion. In: Darji, P.H. (Ed.). Advances in tribology. Intech Open, London, pp. 89–110.10.5772/63657Search in Google Scholar

Azzi, M. and Klemberg-Sapieha, J.-E. (2011). Tribocorrosion test protocols for sliding contacts. In: Landolt, D. and Mischler, S. (Eds.), Tribocorrosion of passive metals and coatings. Woodhead Publishing, Cambridge, pp. 222–238.10.1533/9780857093738.2.222Search in Google Scholar

Azzi, M. and Szpunar, J.A. (2007). Tribo-electrochemical technique for studying tribocorrosion behavior of biomaterials. Biomol. Eng. 24: 443–446, https://doi.org/10.1016/j.bioeng.2007.07.015.Search in Google Scholar PubMed

Azzi, M., Paquette, M., Szpunar, J.A., Klemberg-Sapieha, J.E., and Martinu, L. (2009). Tribocorrosion behaviour of DLC-coated 316L stainless steel. Wear 26: 860–866, https://doi.org/10.1016/j.wear.2009.02.006.Search in Google Scholar

Azzi, M., Amirault, P., Paquette, M., Klemberg-Sapieha, J.E., and Martinu, L. (2010). Corrosion performance and mechanical stability of 316L/DLC coating system: role of interlayers. Surf. Coat. Technol. 204: 3986–3994, https://doi.org/10.1016/j.surfcoat.2010.05.004.Search in Google Scholar

Bailey, R. (2018). Tribocorrosion response of surface-modified Ti in a 0.9 % NaCl solution. Lubricants 6: 1–13, https://doi.org/10.3390/lubricants6040086.Search in Google Scholar

Bailey, R. and Sun, Y. (2018). Corrosion and tribocorrosion performance of pack - carburized commercially pure titanium with limited oxygen diffusion in a 0. 9 % NaCl Solution. J. Bio Tribo Corrosion 4: 1–12, https://doi.org/10.1007/s40735-017-0123-y.Search in Google Scholar

Balamurugan, A., Rajeswari, S., Balossier, G., Rebelo, A.H.S., and Ferreira, J.M.F. (2008). Corrosion aspects of metallic implants – an overview. Mater. Corros. 59: 855–869, https://doi.org/10.1002/maco.200804173.Search in Google Scholar

Bao, M., Wang, X., Yang, L., Qin, G., and Zhang, E. (2018). Tribocorrosion behavior of Ti – Cu Alloy in Hank’s solution for biomedical application. J. Bio Tribo Corrosion 4: 1–13, https://doi.org/10.1007/s40735-018-0142-3.Search in Google Scholar

Bartolomeu, F., Buciumeanu, M., Costa, M.M., Alves, N., Gasik, M., Silva, F.S., and Mirandaa, G. (2019). Multi-material Ti6Al4V & PEEK cellular structures produced by selective laser melting and hot pressing: a tribocorrosion study targeting orthopedic applications. J. Mech. Behav. Biomed. Mater. 89: 54–64, https://doi.org/10.1016/j.jmbbm.2018.09.009.Search in Google Scholar PubMed

Bhat, S.V. (2002). Biomaterials. Springer Dordrecht, Switzerland.10.1007/978-94-010-0328-5Search in Google Scholar

Bidhendi, H.R.A. and Pouranvari, M. (2011). Corrosion study of metallic biomaterials in simulated body fluid. Metalurgija J. Metall. 17: 13–22, https://doi.org/10.30544/384.Search in Google Scholar

Blackwood, D.J. (2003). Biomaterials: past successes and future problems. Corros. Rev. 21: 97–124, https://doi.org/10.1515/corrrev.2003.21.2-3.97.Search in Google Scholar

Bower, K., Murray, S., Reinhart, A., and Nieto, A. (2020). Corrosion resistance of selective laser melted Ti–6Al–4V alloy in salt fog environment. Results Mater. 8: 100122, https://doi.org/10.1016/j.rinma.2020.100122.Search in Google Scholar

Buchanan, R.A., Rigney, E.D.Jr., and Williams, J.M. (1987). Ion implantation of surgical Ti‐6Al‐4V for improved resistance to wear‐accelerated corrosion. J. Biomed. Mater. Res. 21: 55–366, https://doi.org/10.1002/jbm.820210308.Search in Google Scholar PubMed

Buciumeanu, M., Araujo, A., Carvalho, O., Miranda, G., Souza, J.C.M., Silva, F.S., and Henriques, B. (2017). Study of the tribocorrosion behaviour of Ti6Al4V – HA biocomposites. Tribol. Int. 107: 77–84, https://doi.org/10.1016/j.triboint.2016.11.029.Search in Google Scholar

Chandramohan, P., Bhero, S., Obadele, B.A., and Olubambi, P.A. (2017). Laser additive manufactured Ti–6Al–4V alloy: tribology and corrosion studies. Int. J. Adv. Manuf. Technol. 92: 3051–3061, https://doi.org/10.1007/s00170-017-0410-2.Search in Google Scholar

Chauhan, S.R. and Dass, K. (2013). Dry sliding wear behaviour of titanium (Grade 5) alloy by using response surface methodology. Adv. Tribol. 2013: 1–9, https://doi.org/10.1155/2013/272106.Search in Google Scholar

Chen, L.Y., Huang, J.C., Lin, C.H., Pan, C.T., Chen, S.Y., Yang, T.L., Lin, D.Y., Lin, H.K., and Jang, J.S.C. (2017). Anisotropic response of Ti-6Al-4V alloy fabricated by 3D printing selective laser melting. Mater. Sci. Eng., A 682: 389–395, https://doi.org/10.1016/j.msea.2016.11.061.Search in Google Scholar

Chen, Y., Zhang, W., Maitz, M.F., Chen, M., Zhang, H., Mao, J., Zhao, Y., Huang, N., and Wan, G. (2016). Comparative corrosion behavior of Zn with Fe and Mg in the course of immersion degradation in phosphate buffered saline. Corros. Sci. 111: 541–555, https://doi.org/10.1016/j.corsci.2016.05.039.Search in Google Scholar

Cheng, K.Y., Gopal, V., McNallan, M., Manivasagam, G., and Mathew, T.M. (2019). Enhanced tribocorrosion resistance of hard ceramic coated Ti-6Al-4V alloy for hip implant application: in-vitro simulation study. ACS Biomater. Sci. Eng. 5: 4817–4824, https://doi.org/10.1021/acsbiomaterials.9b00609.Search in Google Scholar PubMed

Chiu, T.M., Mahmoudi, M., Dai, W., Elwany, A., Liang, H., and Castaneda, H. (2018). Corrosion assessment of Ti-6Al-4V fabricated using laser powder-bed fusion additive manufacturing. Electrochim. Acta 279: 143–151, https://doi.org/10.1016/j.electacta.2018.04.189.Search in Google Scholar

Ciolac, S., Vasilescu, E., Drob, P., Popa, M.V., and Anghel, M. (2000). Long term in vitro investigation of titanium and some titanium alloys used for surgery implants. Rev. Chem. 51: 36–41.Search in Google Scholar

Contu, F., Elsener, B., and Böhni, H. (2005). Corrosion behaviour of CoCrMo implant alloy during fretting in bovine serum. Corros. Sci. 47: 1863–1875, https://doi.org/10.1016/j.corsci.2004.09.003.Search in Google Scholar

Corbett, R.A. (2003). Proceedings of the Materials and Processes for Medical Devices Conference, September 8–10, 2003: Laboratory Corrosion Testing of Medical Implants. Corrosion Testing Laboratories, Inc., 60 Blue Hen Drive, Newark.Search in Google Scholar

Cunha, L., Andritschky, M., Rebouta, L., and Silva, R. (1998). Corrosion of TiN, (TiAl)N and CrN hard coatings produced by magnetron sputtering. Thin Solid Films 317: 351–355, https://doi.org/10.1016/s0040-6090(97)00624-x.Search in Google Scholar

Czechowska, J., Zima, A., Siek, D., and Ślósarczyk, A. (2016). The importance of chitosan and nano-TiHA in cement-type composites on the basis of calcium sulfate. Ceram. Int. 42: 15559–15567, https://doi.org/10.1016/j.ceramint.2016.07.003.Search in Google Scholar

Dai, N., Zhang, J., Chen, Y., and Zhang, L.C. (2017). Heat treatment degrading the corrosion resistance of selective laser melted Ti-6Al-4V alloy. J. Electrochem. Soc. 164: C428–C434, https://doi.org/10.1149/2.1481707jes.Search in Google Scholar

Deakin, J., Dong, Z., Lynch, B., and Newman, R.C. (2004). De-alloying of type 316 stainless steel in hot, concentrated sodium hydroxide solution. Corros. Sci. 46: 2117–2133, https://doi.org/10.1016/j.corsci.2004.01.011.Search in Google Scholar

DebRoy, T., Wei, H., Zuback, J.S., Mukherjee, T., Elmer, J.W., Milewski, J.O., Beese, A.M., Wilson-Heid, A., De, A., and Zhang, W. (2018). Additive manufacturing of metallic components – process, structure and properties. Prog. Mater. Sci. 92: 112–224, https://doi.org/10.1016/j.pmatsci.2017.10.001.Search in Google Scholar

De Micheli, S.M. and Riesgo, O. (1982). Electrochemical study of corrosion in NiCr dental alloys. Biomaterials 3: 209–212, https://doi.org/10.1016/0142-9612(82)90021-7.Search in Google Scholar PubMed

Dimah, M.K., Albeza, D.F., Borrás, V.A., and Muñoz, A.I. (2012). Study of the biotribocorrosion behaviour of titanium biomedical alloys in simulated body fluids by electrochemical techniques. Wear 294–295: 409–418, https://doi.org/10.1016/j.wear.2012.04.014.Search in Google Scholar

Dlugon, E., Simka, W., Fraczek-Szczypta, A., Niemiec, W., Markowsk, J., Szymanska, M., and Blazewicz, M. (2015). Carbon nanotube-based coatings on titanium. Bull. Mater. Sci. 38: 1339–1344, https://doi.org/10.1007/s12034-015-1019-4.Search in Google Scholar

Elias, C.N., Lima, J.H.C., Valiev, R., and Meyers, M.A. (2008). Biomedical applications of titanium and its alloys. J. Miner. Met. Mater. Soc. 60: 46–49, https://doi.org/10.1007/s11837-008-0031-1.Search in Google Scholar

Eliaz, N. (2019). Corrosion of metallic biomaterials: a review. Materials 12: 1–91, https://doi.org/10.3390/ma12030407.Search in Google Scholar PubMed PubMed Central

Eliaz, N. and Hakshur, K. (2012). Fundamentals of tribology and the use of ferrography and bio-ferrography for monitoring the degradation of natural and artificial joints. In: Eliaz, N. (Ed.). Degradation of implant materials. Springer, New York, NY, pp. 253–302.10.1007/978-1-4614-3942-4_10Search in Google Scholar

Escobar, D.H., Champagne, S., Yilmazer, H., Dikici, B., Boehlert, C.J., and Hermawan, H. (2019). Current status and perspectives of zinc-based absorbable alloys for biomedical applications. Acta Biomater. 97: 1–22, https://doi.org/10.1016/j.actbio.2019.07.034.Search in Google Scholar PubMed

Ettefagh, A.H., Zeng, C., Guo, S., and Raush, J. (2019). Corrosion behavior of additively manufactured Ti-6Al-4V parts and the effect of post annealing. Addit. Manuf. 28: 252–258, https://doi.org/10.1016/j.addma.2019.05.011.Search in Google Scholar

Farè, S., Lecis, N., Vedani, M., Silipigni, A., and Favoino, P. (2012). Properties of nitrided layers formed during plasma nitriding of commercially pure Ti and Ti-6Al-4V alloy. Surf. Coat. Technol. 206: 2287–2292, https://doi.org/10.1016/j.surfcoat.2011.10.006.Search in Google Scholar

Ferreira, D.F., Almeida, S.M.A., Soares, R.B., Juliani, L., Bracarense, A.Q., Lins, V.F.C., and Junqueira, R.M.R. (2018). Synergism between mechanical wear and corrosion on tribocorrosion of a titanium alloy in a Ringer solution. J. Mater. Res. Technol. 8: 1593–1600, https://doi.org/10.1016/j.jmrt.2018.11.004.Search in Google Scholar

Fontana, M.G. (1980). Corrosion engineering and corrosion science, 3rd ed. McGraw Hill, New York.Search in Google Scholar

Francis, A., Yang, Y., Virtanen, S., and Boccaccini, A.R. (2015). Iron and iron-based alloys for temporary cardiovascular applications. J. Mater. Sci.: Mater. Med. 26: 1–16, https://doi.org/10.1007/s10856-015-5473-8.Search in Google Scholar PubMed

Gao, A., Hang, R., Bai, L., Tang, B., and Chu, P.K. (2018). Electrochemical surface engineering of titanium-based alloys for biomedical application. Electrochim. Acta 271: 699–718, https://doi.org/10.1016/j.electacta.2018.03.180.Search in Google Scholar

García, A.J. (2011). Surface modification of biomaterials. In: Atala, A., Lanza, R., Thomson, J. and Nerem, R. (Eds.), Principles of regenerative medicine, 2nd ed. Elsevier, Amsterdam, Netherlands.10.1016/B978-0-12-381422-7.10036-7Search in Google Scholar

García, E., Louvier-Hernández, J.F., Mendoza-Leal, G., Flores-Martínez, M., and Hernández-Navarro, C. (2020). Tribological study of HAp/CTS coatings produced by electrodeposition process on 316L stainless steel. Mater. Lett. 277: 12836, https://doi.org/10.1016/j.matlet.2020.128336.Search in Google Scholar

Gayathri, B. and Bhupathi, S. (2019). Investigation of corrosion protection performance of magnesium incorporated hydroxyapatite coating on surgical grade titanium. Mater. Today: Proc. 18: 1678–1685, https://doi.org/10.1016/j.matpr.2019.05.264.Search in Google Scholar

Ghosh, M., Hartmann, H., Jakobi, M., März, L., Bichmann, L., Freudenmann, L.K., Mühlenbruch, L., Segan, S., Rammense, H.G., Schneiderhan-Marra, N., et al.. (2020). The impact of biomaterial cell contact on the immunopeptidome. Front. Bioeng. Biotechnol. 8: 1–13, https://doi.org/10.3389/fbioe.2020.571294.Search in Google Scholar PubMed PubMed Central

Gilbert, J.L. and Mali, S.A. (2012). Medical implant corrosion: electrochemistry at metallic biomaterial surfaces. In: Eliaz, N. (Ed.). Degradation of implant materials. Springer, New York, NY, pp. 1–28.10.1007/978-1-4614-3942-4_1Search in Google Scholar

Gorejová, R., Oriňaková, R., Králová, Z.O., Sopčák, T., Šišoláková, I., Schnitzer, M., Kohan, M., and Hudák, R. (2023). Electrochemical deposition of a hydroxyapatite layer onto the surface of porous additively manufactured Ti6Al4V scaffolds. Surf. Coat. Technol. 455: 129207, https://doi.org/10.1016/j.surfcoat.2022.129207.Search in Google Scholar

Gopal, V. and Manivasagam, G. (2020). Wear – corrosion synergistic effect on Ti-6Al-4V alloy in H2O2 and albumin environment. J. Alloys Compd. 830: 154539, https://doi.org/10.1016/j.jallcom.2020.154539.Search in Google Scholar

Gopal, V., Chandran, M., Rao, M.S.R., Mischler, S., Cao, S., and Manivasagam, G. (2017). Tribocorrosion and electrochemical behaviour of nanocrystalline diamond coated Ti based alloys for orthopaedic application. Tribol. Int. 106: 88–100, https://doi.org/10.1016/j.triboint.2016.10.040.Search in Google Scholar

Gopi, D., Karthika, A., Rajeswari, D., Kavitha, L., Pramod, V., and Dwivedi, J. (2014). Investigation on corrosion protection and mechanical performance of minerals substituted hydroxyapatite coating on HELCDEB-treated titanium using pulsed electrodeposition method. RSC Adv. 4: 34751–34759, https://doi.org/10.1039/c4ra04484c.Search in Google Scholar

Guo, Z., Panga, X., Yan, Y., Gao, K., Volinsky, A.A., and Zhang, T.Y. (2015). CoCrMo alloy for orthopedic implant application enhanced corrosion and tribocorrosion properties by nitrogen ion implantation. Appl. Surf. Sci. 347: 23–34, https://doi.org/10.1016/j.apsusc.2015.04.054.Search in Google Scholar

Habibovic, P. and Barralet, J.E. (2011). Bioinorganics and biomaterials: bone repair. Acta Biomater. 7: 3013–3026, https://doi.org/10.1016/j.actbio.2011.03.027.Search in Google Scholar PubMed

Hameed, P., Gopal, V., Bjorklund, S., Ganvir, A., Sen, D., Markocsan, N., and Manivasagam, G. (2019). Axial suspension plasma apraying: an ultimate technique to tailor Ti6Al4V surface with HAp for orthopaedic applications. Colloids Surf. B Biointerfaces 173: 806–815, https://doi.org/10.1016/j.colsurfb.2018.10.071.Search in Google Scholar PubMed

Hammood, A.S., Thair, L., Altawaly, H.D., and Parvin, N. (2019). Tribocorrosion behaviour of Ti – 6Al – 4V alloy in biomedical implants: effects of applied load and surface roughness on material degradation. J. Bio Tribo Corrosion 5: 1–12, https://doi.org/10.1007/s40735-019-0277-x.Search in Google Scholar

Hanawa, T. (2012). Degradation of dental implants. In: Eliaz, N. (Ed.). Degradation of implant materials. Springer, New York, NY, pp. 57–78.10.1007/978-1-4614-3942-4_3Search in Google Scholar

Heinl, P., Körner, C., and Singer, R.F. (2008). Selective electron beam melting of cellular titanium: mechanical properties. Adv. Eng. Mater. 10: 882–888, https://doi.org/10.1002/adem.200800137.Search in Google Scholar

Hryniewicz, T., Rokicki, R., and Rokosz, K. (2009). Corrosion and surface characterization of titanium biomaterial after magnetoelectropolishing. Surf. Coat. Technol. 203: 1508–1515, https://doi.org/10.1016/j.surfcoat.2008.11.028.Search in Google Scholar

Huang, H.H., Chiu, Y.H., Lee, T.H., Wu, S.C., Yang, H.W., Su, K.H., and Hsu, C.C. (2003). Ion release from NiTi orthodontic wires in artificial saliva with various acidities. Biomaterials 24: 3585–3592, https://doi.org/10.1016/s0142-9612(03)00188-1.Search in Google Scholar PubMed

Huang, W., Wang, Z., Liu, C., and Yu, Y. (2015). Wear and electrochemical corrosion behavior of biomedical Ti–25Nb–3Mo–3Zr–2Sn alloy in simulated physiological solutions. J. Bio Tribo Corrosion 1: 107780, https://doi.org/10.1007/s40735-014-0001-9.Search in Google Scholar

Huang, X., Liu, H., Wang, Z., Qiao, L., Su, Y., and Yan, Y. (2022). Effect of surface oxidation on wear and tribocorrosion behavior of forged and selective laser melting-based TC4 alloys. Tribol. Int. 174: 107780, https://doi.org/10.1016/j.triboint.2022.107780.Search in Google Scholar

Hui, T., Kubacki, G.W., and Gilbert, J.L. (2014). Voltage and wear debris from Ti-6Al-4V interact to affect cell viability during in-vitro fretting corrosion. J. Biomed. Mater. Res., Part A 106: 160–167, https://doi.org/10.1002/jbm.a.36220.Search in Google Scholar PubMed

Kato, K. (2002). Proceedings of the Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology: Classification of Wear Mechanisms/Models. Institution of Mechanical Engineers, London.10.1243/135065002762355280Search in Google Scholar

Kaur, S., Sharma, S., and Bala, N. (2019). A comparative study of corrosion resistance of biocompatible coating on titanium alloy and stainless steel. Mater. Chem. Phys. 238: 121923, https://doi.org/10.1016/j.matchemphys.2019.121923.Search in Google Scholar

Keselowsky, B.G., Wang, L., Schwartz, Z., Garcia, A.J., and Boyan, B.D. (2007). Integrin α5 controls osteoblastic proliferation and differentiation responses to titanium substrates presenting different roughness characteristics in a roughness independent manner. J. Biomed. Mater. Res., Part A 80: 700–710, https://doi.org/10.1002/jbm.a.30898.Search in Google Scholar PubMed

Kim, G.E. and Walker, J. (2007). Successful application of nanostructured titanium dioxide coating for high-pressure acid-leach application. J. Therm. Spray Technol. 16: 34–39, https://doi.org/10.1007/s11666-006-9004-5.Search in Google Scholar

Koike, M. and Fujii, H. (2001). The corrosion resistance of pure titanium in organic acids. Biomaterials 22: 2931–2936, https://doi.org/10.1016/s0142-9612(01)00040-0.Search in Google Scholar PubMed

Kong, D., Dong, C., Ni, X., and Li, X. (2019). Corrosion of metallic materials fabricated by selective laser melting. npj Mater. Degrad. 3: 1–14, https://doi.org/10.1038/s41529-019-0086-1.Search in Google Scholar

Kuhn, A.T. (1981). Corrosion of Co-Cr alloys in aqueous environments. Biomaterials 2: 68–77, https://doi.org/10.1016/0142-9612(81)90002-8.Search in Google Scholar PubMed

Kuhn, A.T., Neufeld, P., and Rae, T. (2009). Synthetic environments for the testing of metallic biomaterials. ASTM Int. 79: 1–300.10.1520/STP26001SSearch in Google Scholar

Kumar, D.D., Kaliaraj, G.S., Kirubaharan, A.M.K., AlagarsamY, K., Vishwakarma, V., and Baskaran, R. (2019). Biocorrosion and biological properties of sputtered ceramic carbide coatings for biomedical applications. Surf. Coat. Technol. 374: 569–578, https://doi.org/10.1016/j.surfcoat.2019.06.022.Search in Google Scholar

Kumari, A., Yadav, S.K., and Yadav, S.C. (2010). Biodegradable polymeric nanoparticles based drug delivery systems. Colloids Surf. B Biointerfaces 75: 1–18, https://doi.org/10.1016/j.colsurfb.2009.09.001.Search in Google Scholar PubMed

Laing, P.G. (1973). Compatibility of biomaterials. Orthop. Clin. N. Am. 4: 249–273, https://doi.org/10.1016/s0030-5898(20)30792-6.Search in Google Scholar

Landolt, D., Mischler, S., and Stemp, M. (2001). Electrochemical methods in tribocorrosion: a critical appraisal. Electrochim. Acta 46: 3913–3929, https://doi.org/10.1016/s0013-4686(01)00679-x.Search in Google Scholar

Leon, A. and Aghion, E. (2017). Effect of surface roughness on corrosion fatigue performance of AlSi10Mg alloy produced by selective laser melting (SLM). Mater. Charact. 131: 188–194, https://doi.org/10.1016/j.matchar.2017.06.029.Search in Google Scholar

Leon, A., Levy, G.K., Ron, T., Shirizly, A., and Aghion, E. (2020). The effect of strain rate on stress corrosion performance of Ti6Al4V alloy produced by additive manufacturing process. J. Mater. Res. Technol. 9: 4097–4105, https://doi.org/10.1016/j.jmrt.2020.02.035.Search in Google Scholar

Li, D.G., Wang, J.D., Chen, D.R., and Liang, P. (2015). Influence of molybdenum on tribo-corrosion behavior of 316L stainless steel in artificial aaliva. J. Bio Tribo Corrosion 1: 1–9.10.1007/s40735-015-0014-zSearch in Google Scholar

Li, Y., Zhou, J., Pavanram, P., Leeflang, M.A., Fockaert, L.I., Pouran, B., Tümer, N., Schröder, K.U., Mol, J.M.C., Weinans, H., et al.. (2018). Additively manufactured biodegradable porous magnesium. Acta Biomater. 67: 378–392, https://doi.org/10.1016/j.actbio.2017.12.008.Search in Google Scholar PubMed

Li, Y., Jahr, H., Zhou, J., and Zadpoo, A.A. (2020). Additively manufactured biodegradable porous metals. Acta Biomater. 115: 29–50, https://doi.org/10.1016/j.actbio.2020.08.018.Search in Google Scholar PubMed

Licausi, M.P. (2017). Analysis of tribocorrosion behavior of biomedical powder metallurgy titanium alloys, Ph. D Thesis. Valencia Polytechnic University, Valencia.Search in Google Scholar

Licausi, M.P., Muñoz, A.I., Borrás, V.A., and Espallargas, N. (2015). Tribocorrosion mechanisms of Ti6Al4V in artificial saliva by zero-resistance ammetry (ZRA) technique. J. Bio Tribo Corrosion 1: 1–11, https://doi.org/10.1007/s40735-015-0008-x.Search in Google Scholar

López-Ortega, A., Arana, J.L., and Bayón, R. (2018). Tribocorrosion of passive materials: a review on test procedures and standards. Int. J. Corrosion 2018: 1–24, https://doi.org/10.1155/2018/7345346.Search in Google Scholar

Love, C.A., Cook, R.B., Harvey, T.J., Dearnley, P.A., and Wood, R.J.K. (2013). Diamond like carbon coatings for potential application in biological implants – a review. Tribol. Int. 63: 141–150, https://doi.org/10.1016/j.triboint.2012.09.006.Search in Google Scholar

Manhabosco, T.M. and Müller, I.L. (2009). Tribocorrosion of diamond-like carbon deposited on Ti6Al4V. Tribol. Lett. 33: 193–197, https://doi.org/10.1007/s11249-009-9408-8.Search in Google Scholar

Manivasagam, G., Singh, A.K., Asokamani, R., and Gogia, A.K. (2009). Ti based biomaterials, the ultimate choice for orthopaedic implants. A review. Prog. Mater. Sci. 54: 397–425, https://doi.org/10.1016/j.pmatsci.2008.06.004.Search in Google Scholar

Manivasagam, G., Dhinasekaran, D., and Rajamanickam, A. (2010). Biomedical implants: corrosion and its prevention – a review. Recent Pat. Corros. Sci. 2: 40–54, https://doi.org/10.2174/1877610801002010040.Search in Google Scholar

Manjunath, S.V.M., Mohan, R.L., and Bera, P. (2017). Corrosion and wear properties of Ti/tetrahedral amorphous carbon multilayered coating. J. Bio Tribo Corrosion 3: 1–10.10.1007/s40735-017-0100-5Search in Google Scholar

Manoj, A., Kasar, A.K., and Menezes, P.L. (2019). Tribocorrosion of porous titanium used in biomedical applications. J. Bio Tribo Corrosion 5: 1–16, https://doi.org/10.1007/s40735-018-0194-4.Search in Google Scholar

Marie, P. (1992). Physiology of bone tissue. Immuno-Anal. Biol. Specialisee 7: 17–24, https://doi.org/10.1016/s0923-2532(05)80182-6.Search in Google Scholar

Mathew, M.T. and Wimmer, M.A. (2011). Tribocorrosion in artificial joints: in vitro testing and clinical implications. In: Yu, Y. (Ed.). Bio-tribocorrosion in biomaterials and medical implants. Woodhead Publishing, Swaston, UK, pp. 341–371.10.1533/9780857098603.3.341Search in Google Scholar

Mathew, M.T., Pai, P.S., Pourzal, R., Fischer, A., and Wimmer, M.A. (2009). Significance of tribocorrosion in biomedical applications: overview and current status. Adv. Tribol. 2009: 1–12, https://doi.org/10.1155/2009/250986.Search in Google Scholar

Mathew, M.T., Nagelli, C., Pourzal, R., Fischer, A., Laurent, M.P., Jacobs, J.J., and Wimmer, M.A. (2014). Tribolayer formation in a metal-on-metal (MoM) hip joint: an electrochemical investigation. J. Mech. Behav. Biomed. Mater. 29: 199–212, https://doi.org/10.1016/j.jmbbm.2013.08.018.Search in Google Scholar PubMed PubMed Central

Maurer, A.M., Brown, S.A., Payer, J.H., Merritt, K., and Kawalec, J.S. (1993). Reduction of fretting corrosion of Ti-6Al-4V by various surface treatments. J. Orthop. Res. 11: 865–873, https://doi.org/10.1002/jor.1100110613.Search in Google Scholar PubMed

Mears, D.C. (1979). Materials and orthopaedic surgery. Williams & Wilkins, Philadelphia, USA.Search in Google Scholar

Mei, D., Lamaka, S.V., Lu, X., and Zheludkevich, M.L. (2020). Selecting medium for corrosion testing of bioabsorbable magnesium and other metals – a critical review. Corros. Sci. 171: 108722, https://doi.org/10.1016/j.corsci.2020.108722.Search in Google Scholar

Metikoš-Huković, M., Pilić, Z., Babić, R., and Omanović, D. (2006). Influence of alloying elements on the corrosion stability of CoCrMo implant alloy in Hank’s solution. Acta Biomater. 2: 693–700, https://doi.org/10.1016/j.actbio.2006.06.002.Search in Google Scholar PubMed

Mischler, S., Spiegel, A., and Landolt, D. (1999). The role of passive oxide films on the degradation of steel in tribocorrosion systems. Wear 225–229: 1078–1087, https://doi.org/10.1016/s0043-1648(99)00056-3.Search in Google Scholar

Mo, X., Qian, J., Chen, Y., Zhang, W., Xian, P., Tang, S., Zhou, C., Huang, N., Ji, H., Luo, E., et al.. (2021). Corrosion and degradation decelerating alendronate embedded zinc phosphate hybrid coating on biodegradable Zn biomaterials. Corros. Sci. 184: 109398, https://doi.org/10.1016/j.corsci.2021.109398.Search in Google Scholar

Mohan, L., Dhinasekaran, D., Manivasagam, G., Sankaranarayanan, T.S.N., and Asokamani, R. (2012). Electrophoretic deposition of nanocomposite (HAp + TiO 2) on titanium alloy for biomedical applications. Ceram. Int. 38: 3435–3443, https://doi.org/10.1016/j.ceramint.2011.12.056.Search in Google Scholar

Mohazzab, B.F., Jaleh, B., Fattah-Alhosseini, A.F., Mahmoudi, F., and Momeni, A. (2020). Laser surface treatment of pure titanium: microstructural analysis, wear properties, and corrosion behavior of titanium carbide coatings in Hank’s physiological solution. Surface. Interfac. 20: 100597, https://doi.org/10.1016/j.surfin.2020.100597.Search in Google Scholar

Morris, D., Mamidi, S.K., Kamat, S., Cheng, K., Bijukumar, D., Tsai, P., Wu, M.H., Orias, A.A.E., and Mathew, T.M. (2021). Mechanical, electrochemical and biological behavior of 3D printed-porous titanium for biomedical applications. J. Bio Tribo Corrosion 7: 1–15, https://doi.org/10.1007/s40735-020-00457-5.Search in Google Scholar

Movassagh-Alanagh, F., Abdollah-zadeh, A., Asgari, M., and Ghaffari, M.A. (2018). Influence of Si content on the wettability and corrosion resistance of nanocomposite TiSiN films deposited by pulsed-DC PACVD. J. Alloys Compd. 730: 780–792, https://doi.org/10.1016/j.jallcom.2017.12.235.Search in Google Scholar

Munir, K.S., Kingshott, P., and Wen, C. (2015). Carbon nanotube reinforced titanium metal matrix composites prepared by powder metallurgy – a review. Crit. Rev. Solid State Mater. Sci. 40: 38–55, https://doi.org/10.1080/10408436.2014.929521.Search in Google Scholar

Murphy, W., Black, J., and Hastings, G. (2016). Handbook of biomaterial properties, 2nd ed. Springer, New York.10.1007/978-1-4939-3305-1Search in Google Scholar

Nasab, M.B. and Hassan, M.R. (2009). Metallic biomaterials of knee and hip. A review. Trends Biomater. Artif. Organs 24: 69–82.Search in Google Scholar

Nascimento, C.A., Barbosa, J.A., Montalli, V.A.M., Micheletti, F., Milani, R., Pereira, V., Caldeira, L., and Basting, R.T. (2020). Corrosion and micromorphological analysis of temporary stainless steel and titanium alloy anchorage devices. J. Bio Tribo Corrosion 6: 1–8.10.1007/s40735-020-00358-7Search in Google Scholar

Ng, K.W. and Man, H.C. (2012). Laser surface modification of nickel-titanium (NiTi) alloy biomaterials to improve biocompatibility and corrosion resistance. In: Kwok, C.T. (Ed.). Laser surface modification of alloys for corrosion and erosion resistance. Woodhead Publishing, Swaston, UK, pp. 124–151.10.1533/9780857095831.1.124Search in Google Scholar

Okazaki, Y. (2002). Effect of friction on anodic polarization properties of metallic biomaterials. Biomaterials 23: 2071–2077, https://doi.org/10.1016/s0142-9612(01)00337-4.Search in Google Scholar PubMed

Oláh, N., Fogarassy, Z., Furkó, M., Balázsi, C., and Balázsi, K. (2015). Sputtered nanocrystalline ceramic TiC/amorphous C thin films as potential materials for medical applications. Ceram. Int. 41: 5863–5871, https://doi.org/10.1016/j.ceramint.2015.01.017.Search in Google Scholar

Ong, K.L., Lovald, S., and Black, J. (2017). Orthopaedic biomaterials in research and practice, 2nd ed. CRC Press, Florida.Search in Google Scholar

Pan, J., Thierry, D., and Leygraf, C. (1996). Electrochemical impedance spectroscopy study of the passive oxide film on titanium for implant application. Electrochim. Acta 41: 1143–1153, https://doi.org/10.1016/0013-4686(95)00465-3.Search in Google Scholar

Paula, L.O., Sene, A.C., Manfroi, L.A., Vieira, A.A., Ramos, M.A.R., Fukumasu, N.K., Radi, P.A., and Vieira, L. (2018). Tribo-corrosion and corrosion behaviour of titanium alloys with and without DLC films immersed in synthetic urine. J. Bio Tribo Corrosion 4: 1–12, https://doi.org/10.1007/s40735-018-0166-8.Search in Google Scholar

Pejaković, V., Totolin, V., and Rodríguez Ripoll, M. (2018). Tribocorrosion behaviour of Ti6Al4V in artificial seawater at low contact pressures. Tribol. Int. 119: 55–65, https://doi.org/10.1016/j.triboint.2017.10.025.Search in Google Scholar

Ponnusamy, S.G.S. and Muthamizhchelvan, L.M.C. (2018). In vitro corrosion behaviour of Ti – 6Al – 4V and 316L stainless steel alloys for biomedical implant applications. J. Bio Tribo Corrosion 4: 1–8, https://doi.org/10.1007/s40735-017-0118-8.Search in Google Scholar

Ponthiaux, P., Wenger, F., Drees, D., and Celis, J.P. (2004). Electrochemical techniques for studying tribocorrosion processes. Wear 256: 459–468, https://doi.org/10.1016/s0043-1648(03)00556-8.Search in Google Scholar

Pound, B.G. (2014). Corrosion behavior of metallic materials in biomedical applications. I. Ti and its alloys. Corros. Rev. 32: 1–20, https://doi.org/10.1515/corrrev-2014-0007.Search in Google Scholar

Prabakaran, V. (2019). Characterization and corrosion studies of TiAlN PVD coating by using the polarization test method. J. Bio Tribo Corrosion 5: 1–7.10.1007/s40735-019-0220-1Search in Google Scholar

Priyadarshini, B. and Vijayalakshmi, U. (2021). In vitro bioactivity, biocompatibility and corrosion resistance of multi-ionic (Ce/Si) co-doped hydroxyapatite porous coating on Ti-6Al-4 V for bone regeneration applications. Mater. Sci. Eng. C 119: 111620, https://doi.org/10.1016/j.msec.2020.111620.Search in Google Scholar PubMed

Raja, V.S. and Shoji, T. (2007). Stress corrosion cracking. Theory and practice. Woodhead Publishing Limited, Swaston, UK.Search in Google Scholar

Ratner, B., Hoffman, A., Schoen, F., and Lemons, J. (2004). Biomaterials science: an introduction to materials, 2nd ed. Elsevier, Amsterdam.Search in Google Scholar

Ratner, B., Hoffman, A., Schoen, F., and Lemons, J. (2013). Biomaterials science: an introduction to materials, 3rd ed. Elsevier, Amsterdam.Search in Google Scholar

Rautray, T.R., Narayanan, R., and Kim, K.H. (2011). Ion implantation of titanium based biomaterials. Prog. Mater. Sci. 56: 1137–1177, https://doi.org/10.1016/j.pmatsci.2011.03.002.Search in Google Scholar

Revilla, R.I., Liang, J., Godet, S., and Graeve, I.D. (2017). Local corrosion behavior of additive manufactured AlSiMg alloy assessed by SEM and SKPFM. J. Electrochem. Soc. 164: C27–C35, https://doi.org/10.1149/2.0461702jes.Search in Google Scholar

Richard, C., Kowandy, C., Landoulsi, J., Geetha, M., and Ramasawmy, H. (2010). Corrosion and wear behavior of thermally sprayed nano ceramic coatings on commercially pure titanium and Ti-13Nb-13Zr substrates. Int. J. Refract. Metals Hard Mater. 28: 115–123, https://doi.org/10.1016/j.ijrmhm.2009.08.006.Search in Google Scholar

Rossi, S., Deflorian, F., Zen, M., and Fedrizzi, L. (2000). Wear-corrosion of nitrided steel: corrosion potential monitoring to evaluate the effect of test parameters. Mater. Corros. 51: 552–556, https://doi.org/10.1002/1521-4176(200008)51:8<552::aid-maco552>3.0.co;2-g.10.1002/1521-4176(200008)51:8<552::AID-MACO552>3.0.CO;2-GSearch in Google Scholar

Rylska, D., Sokołowski, G., Sokołowski, J., and Łukomska-Szymańska, M. (2017). Chemical passivation as a method of improving the electrochemical corrosion resistance of Co-Cr-based dental alloy. Acta Bioeng. Biomech. 19: 73–78.Search in Google Scholar

Saikko, V. (1998). A multidirectional motion pin-on-disk wear test method for prosthetic joint materials. J. Biomed. Mater. Res. 41: 58–64, https://doi.org/10.1002/(sici)1097-4636(199807)41:1<58::aid-jbm7>3.0.co;2-p.10.1002/(SICI)1097-4636(199807)41:1<58::AID-JBM7>3.3.CO;2-USearch in Google Scholar

Sampaio, M., Buciumeanu, M., Henriques, B., Silva, F.S., Souza, J.C.M., and Gomes, J.R. (2016). Tribocorrosion behavior of veneering biomedical PEEK to Ti6Al4V structures. J. Mech. Behav. Biomed. Mater. 54: 123–130, https://doi.org/10.1016/j.jmbbm.2015.09.010.Search in Google Scholar PubMed

Sasikumar, Y. and Rajendran, N. (2018). Effect of fluoride concentration and pH on corrosion behavior of Ti–15Mo in artificial saliva. J. Bio Tribo Corrosion 4: 1–8, https://doi.org/10.1007/s40735-017-0119-7.Search in Google Scholar

Shao, T., Ge, F., Pei, C., Huang, F., Sun, D., and Zhang, S. (2018). Effects of Si content on tribo-corrosion behavior of Cr1-xSixN coatings prepared via magnetron sputtering. Surf. Coat. Technol. 356: 11–18, https://doi.org/10.1016/j.surfcoat.2018.09.031.Search in Google Scholar

Sharma, N., Singh, G., Hegab, H., Mia, M., and Batra, N.K. (2019). Tribo-corrosion characterization of NiTiCu alloy for bio-implant applications. Mater. Res. Express 6: 1–15, https://doi.org/10.1088/2053-1591/ab2d95.Search in Google Scholar

Shuai, C., He, C., Feng, P., Guo, W., Gao, C., Wu, P., Yang, Y., and Bin, S. (2018a). Biodegradation mechanisms of selective laser-melted Mg–xAl–Zn alloy: grain size and intermetallic phase. Virtual Phys. Prototyp. 13: 59–69, https://doi.org/10.1080/17452759.2017.1408918.Search in Google Scholar

Shuai, C., Xue, L., Gao, C., Yang, Y., Peng, S., and Zhang, Y. (2018b). Selective laser melting of Zn–Ag alloys for bone repair: microstructure, mechanical properties and degradation behaviour. Virtual Phys. Prototyp. 13: 146–154, https://doi.org/10.1080/17452759.2018.1458991.Search in Google Scholar

Sin, J.R. (2015). Investigation of the corrosion and tribocorrosion behaviour of metallic biomaterials, PhD thesis. Luleå University of Technology, Luleå, Sweden.Search in Google Scholar

Souza, J.C.M., Tajiri, H.A., Morsch, C.S., Buciumeanu, M., Mathew, M.T., Silva, F.S., and Henriques, B. (2015). Tribocorrosion behavior of Ti6Al4V coated with a bio-absorbable polymer for biomedical applications. J. Bio Tribo Corrosion 1: 2–7, https://doi.org/10.1007/s40735-015-0029-5.Search in Google Scholar

Stachowiak, G.W. and Batchelor, A.W. (2001). Engineering tribology, 2nd ed. Elsevier, Amsterdam, Netherlands.Search in Google Scholar

Stack, M.M., Rodling, J., Mathew, M.T., Jawan, H., Huang, W., Park, G., and Hodge, C. (2010). Micro-abrasion-corrosion of a Co-Cr/UHMWPE couple in Ringer’s solution: an approach to construction of mechanism and synergism maps for application to bio-implants. Wear 269: 376–382, https://doi.org/10.1016/j.wear.2010.04.022.Search in Google Scholar

Stack, M.M., Huang, W., Wang, G., and Hodge, C. (2011). Some views on the construction of bio-tribo-corrosion maps for titanium alloys in Hank’s solution: particle concentration and applied loads effects. Tribol. Int. 44: 1827–1837, https://doi.org/10.1016/j.triboint.2011.07.009.Search in Google Scholar

Sukuroglu, E.E., Totik, Y., Arslan, E., and Efeoglu, I. (2015). Analysis of tribo-corrosion properties of MAO/DLC coatings using a duplex process on Ti 6 Al 4 V alloys. J. Bio Tribo Corrosion 1: 1–13, https://doi.org/10.1007/s40735-015-0022-z.Search in Google Scholar

Sundararajan, T., Mudali, U.K., Nair, K.G.M., Rajeswari, S., and Subbaiyan, M. (1999). In vitro corrosion evaluation of nitrogen ion implanted titanium in simulated body fluid. Mater. Corros. 50: 344–349, https://doi.org/10.1002/(sici)1521-4176(199906)50:6<344::aid-maco344>3.0.co;2-4.10.1002/(SICI)1521-4176(199906)50:6<344::AID-MACO344>3.0.CO;2-4Search in Google Scholar

Surmeneva, M.A., Chudinova, E.A., Chernozem, R.V., Lapanje, A., Koptyug, A.V., Rijavec, T., Loza, K., Pryma, O., Epple, M., Wittmar, A., et al.. (2020). Development of a bone substitute material based on additive manufactured Ti6Al4V alloys modified with bioceramic calcium carbonate coating: characterization and antimicrobial properties. Ceram. Int. 46: 25661–25670, https://doi.org/10.1016/j.ceramint.2020.07.041.Search in Google Scholar

Suzuki, M.K., Martins, D.A., Costa, M.T., Ferreira, A.C., and Ferreira, F.A. (2018). Ions release evaluation and changes in mini-implant orthodontic surface. J. Contemp. Dent. Pract. 19: 910–917, https://doi.org/10.5005/jp-journals-10024-2356.Search in Google Scholar

Take, S., Kikuchi, K., Suda, S., Izawa, S., and Itoi, Y. (2014). Preparation and evaluation of Zn doped HAp plasma spray biocombatible coatings on titanium. ECS Trans. 58: 17–22, https://doi.org/10.1149/05838.0017ecst.Search in Google Scholar

Tamilselvi, S., Raman, V., and Rajendran, N. (2010). Evaluation of corrosion behavior of surface modified Ti–6Al–4V ELI alloy in hanks solution. J. Appl. Electrochem. 40: 285–293, https://doi.org/10.1007/s10800-009-9972-5.Search in Google Scholar

Tiainen, V.M. (2001). Amorphous carbon as a bio-mechanical coating-mechanical properties and biological applications. Diamond Relat. Mater. 10: 153–160, https://doi.org/10.1016/s0925-9635(00)00462-3.Search in Google Scholar

Tohidi, P.M.S., Safavi, M.S., Etminanfar, M., and Khalil-Allafi, J. (2020). Pulsed electrodeposition of compact, corrosion resistant, and bioactive HAp coatings by application of optimized magnetic field. Mater. Chem. Phys. 254: 123511, https://doi.org/10.1016/j.matchemphys.2020.123511.Search in Google Scholar

Totten, G.E. and Liang, H. (2004). Mechanical tribology. CRC Press, Boca Raton.10.1201/9780203970911Search in Google Scholar

Traisnel, M., Maguer, D.L., Hildebrand, H.L., and Lost, A. (1990). Corrosion of surgical implants. Clin. Mater. 5: 309–318, https://doi.org/10.1016/0267-6605(90)90030-y.Search in Google Scholar

Trino, L.D., Bronze-Uhle, E.S., Ramachandran, A., Lisboa-Filho, P.N., Mathew, M.T., and George, A. (2018). Titanium surface bio-functionalization using osteogenic peptides: surface chemistry, biocompatibility, corrosion and tribocorrosion aspects. J. Mech. Behav. Biomed. Mater. 81: 26–38, https://doi.org/10.1016/j.jmbbm.2018.02.024.Search in Google Scholar

Tuomi, J.T., Björkstrand, R.V., Pernu, M.L., Salmi, M.V.J., Huotilainen, E.I., Wolff, J.E.H., Vallittu, P.K., and Mäkitie, A.A. (2017). In vitro cytotoxicity and surface topography evaluation of additive manufacturing titanium implant materials. J. Mater. Sci.: Mater. Med. 28: 1–7, https://doi.org/10.1007/s10856-017-5863-1.Search in Google Scholar PubMed

Venezuela, J. and Dargusch, M.S. (2019). The influence of alloying and fabrication techniques on the mechanical properties, biodegradability and biocompatibility of zinc: a comprehensive review. Acta Biomater. 87: 1–40, https://doi.org/10.1016/j.actbio.2019.01.035.Search in Google Scholar PubMed

Venkatesan, S.P., Jeevahan, J., Purusothaman, M., Venkatesh, S., and Vimal, M.R. (2021). Corrosion and mechanical behavior of plasma nitrated metallic biomaterial surfaces. Mater. Today: Proc. 47: 938–943, https://doi.org/10.1016/j.matpr.2021.04.637.Search in Google Scholar

Vilhena, L., Oppong, G., and Ramalho, A. (2019). Tribocorrosion of different biomaterials under reciprocating sliding conditions in artificial saliva. Lubric. Sci. 31: 364–380, https://doi.org/10.1002/ls.1478.Search in Google Scholar

Virtanen, S. (2008). Corrosion of biomedical implant materials. Corros. Rev. 26: 147–172, https://doi.org/10.1515/corrrev.2008.147.Search in Google Scholar

Virtanen, S. (2012). Degradation of titanium and its alloys. In: Eliaz, N. (Ed.). Degradation of implant materials. Springer, New York, pp. 29–55.10.1007/978-1-4614-3942-4_2Search in Google Scholar

Wang, J.L., Xu, J.K., Hopkins, C., Chow, D.H.K., and Qin, L. (2020). Biodegradable magnesium-based implants in orthopedics: a general review and perspectives. Adv. Sci. 7: 1–19, https://doi.org/10.1002/advs.201902443.Search in Google Scholar PubMed PubMed Central

Wang, X., Shao, X., Dai, T., Xu, F., Zhou, J.G., Qu, G., Tian, L., Liu, B., and Liu, Y. (2019). In vivo study of the efficacy, biosafety, and degradation of a zinc alloy osteosynthesis system. Acta Biomater. 92: 351–361, https://doi.org/10.1016/j.actbio.2019.05.001.Search in Google Scholar PubMed

Wang, Z., Yan, Y., Wu, Y., Huang, X., Zhang, Y., Su, Y., and Qiao, L. (2022). Corrosion and tribocorrosion behavior of equiatomic refractory medium entropy TiZr (Hf, Ta, Nb) alloys in chloride solutions. Corros. Sci. 199: 1–19, https://doi.org/10.1016/j.corsci.2022.110166.Search in Google Scholar

Whitesides, G.M. and Wong, A.P. (2006). The intersection of biology and materials science. MRS Bull. 31: 19–27, https://doi.org/10.1557/mrs2006.2.Search in Google Scholar

Williams, D.F. (Ed.) (1986). Definitions in Biomaterials: Proceedings of a Consensus Conference of the European Society for Biomaterials, March 3–5, 1986. Elsevier, Chester, England.Search in Google Scholar

Williams, J.A. (2006). Engineering tribology. Cambridge University Press, Cambridge.Search in Google Scholar

Wright, K.W.J., Dobbs, H.S., and Scales, J.T. (1982). Wear studies on prosthetic materials using the pin-on-disc machine. Biomaterials 3: 41–48, https://doi.org/10.1016/0142-9612(82)90060-6.Search in Google Scholar PubMed

Xu, X., Lu, Y., Sundberg, K.L., Liang, J., and Sisson, R.D.Jr (2017). Effect of annealing treatments on the microstructure, mechanical properties and corrosion behavior of direct metal laser sintered Ti-6Al-4V. J. Mater. Eng. Perform. 26: 2572–2582, https://doi.org/10.1007/s11665-017-2710-y.Search in Google Scholar

Yan, Y. (2006). Corrosion and tribo-corrosion behaviour of metallic orthopaedic implant, Ph.D. thesis. University of Leeds, Leeds.Search in Google Scholar

Yang, Y., Yuan, F., Gao, C., Feng, P., Xue, L., He, S., and Shuai, C. (2018). A combined strategy to enhance the properties of Zn by laser rapid solidification and laser alloying. J. Mech. Behav. Biomed. Mater. 82: 51–60, https://doi.org/10.1016/j.jmbbm.2018.03.018.Search in Google Scholar PubMed

Yazdi, R., Ghasemi, H.M., Wang, C., and Neville, A. (2017). Bio-corrosion behaviour of oxygen diffusion layer on Ti-6Al-4V during tribocorrosion. Corros. Sci. 128: 23–32, https://doi.org/10.1016/j.corsci.2017.08.031.Search in Google Scholar

Yazdi, R., Ghasemi, H.M., Abedini, M., Wang, C., and Neville, A. (2018). Mechanism of tribofilm formation on Ti6Al4V oxygen diffusion layer in a simulated body fluid. J. Mech. Behav. Biomed. Mater. 77: 660–670, https://doi.org/10.1016/j.jmbbm.2017.10.020.Search in Google Scholar PubMed

Yu, F., Addison, O., and Davenport, A.J. (2015). A synergistic effect of albumin and H2O2 accelerates corrosion of Ti6Al4V. Acta Biomater. 26: 355–365, https://doi.org/10.1016/j.actbio.2015.07.046.Search in Google Scholar PubMed

Yue, T.M., Yu, J.K., Mei, Z., and Man, H.C. (2002). Excimer laser surface treatment of Ti-6Al-4V alloy for corrosion resistance enhancement. Mater. Lett. 52: 206–212, https://doi.org/10.1016/s0167-577x(01)00395-0.Search in Google Scholar

Zhang, L., Duan, Y., Gao, Z., Ma, J., Liu, R., Liu, S., Tu, Z., Liu, Y., Bai, C., Cui, L., et al.. (2016). Graphene enhanced anti-corrosion and biocompatibility of NiTi alloy. NanoImpact 7: 7–14, https://doi.org/10.1016/j.impact.2016.10.003.Search in Google Scholar

Zhang, T.F., Deng, Q.Y., Liu, B., Wu, B.J., Jing, F.J., Leng, Y.X., and Huang, N. (2015). Wear and corrosion properties of diamond like carbon (DLC) coating on stainless steel, CoCrMo and Ti6Al4V substrates. Surf. Coat. Technol. 273: 12–19, https://doi.org/10.1016/j.surfcoat.2015.03.031.Search in Google Scholar

Zhang, Y., Addison, O., Yu, F., Troconis, B.C.R., Scully, J.R., and Davenport, A.J. (2018). Time-dependent enhanced corrosion of Ti6Al4V in the presence of H2O2 and albumin. Sci. Rep. 8: 1–11, https://doi.org/10.1038/s41598-018-21332-x.Search in Google Scholar PubMed PubMed Central

Zhao, B., Wang, H., Qiao, N., Wang, C., and Hu, M. (2017). Corrosion resistance characteristics of a Ti-6Al-4V alloy scaffold that is fabricated by electron beam melting and selective laser melting for implantation in vivo. Mater. Sci. Eng. C 70: 832–841, https://doi.org/10.1016/j.msec.2016.07.045.Search in Google Scholar PubMed

Zhao, G.H., Aune, R.E., and Espallargas, N. (2016a). Tribocorrosion studies of metallic biomaterials: the effect of plasma nitriding and DLC surface modifications. J. Mech. Behav. Biomed. Mater. 63: 100–114, https://doi.org/10.1016/j.jmbbm.2016.06.014.Search in Google Scholar PubMed

Zhao, J., Xu, D., Shahzad, M.B., Kang, Q., Sun, Y., Sun, Z., Zhang, S., Ren, L., Yang, C., and Yang, K. (2016b). Effect of surface passivation on corrosion resistance and antibacterial properties of Cu-bearing 316L stainless steel. Appl. Surf. Sci. 386: 371–380, https://doi.org/10.1016/j.apsusc.2016.06.036.Search in Google Scholar

Zhecheva, A., Sha, W., Malinov, S., and Long, A. (2005). Enhancing the microstructure and properties of titanium alloys through nitriding and other surface engineering methods. Surf. Coat. Technol. 200: 2192–2207, https://doi.org/10.1016/j.surfcoat.2004.07.115.Search in Google Scholar

Zheng, Y.F., Gu, X.N., and Witte, F. (2014). Biodegradable metals. Technical report. Mater. Sci. Eng. R Rep. 77: 1–34, https://doi.org/10.1016/j.mser.2014.01.001.Search in Google Scholar

Zhou, Y., Wu, P., Yang, Y., Gao, D., Feng, P., Gao, C., Wu, H., Liu, Y., Bian, H., and Shuai, C. (2016). The microstructure, mechanical properties and degradation behavior of laser-melted Mg-Sn alloys. J. Alloys Compd. 687: 109–114, https://doi.org/10.1016/j.jallcom.2016.06.068.Search in Google Scholar

Zierold, A.A. (1924). Reaction of bone to various metals. Arch. Surg. 9: 365–412, https://doi.org/10.1001/archsurg.1924.01120080133008.Search in Google Scholar

Zou, J., Wang, Z., Ma, Y., Yan, Y., and Qiao, L. (2022). Role of gradient nano-structured surface in collapsed pitting corrosion on AISI 316L stainless steel during tribocorrosion. Corros. Sci. 197: 110043, https://doi.org/10.1016/j.corsci.2021.110043.Search in Google Scholar

Received: 2023-02-18
Accepted: 2023-08-28
Published Online: 2023-11-13
Published in Print: 2024-02-26

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 6.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2023-0008/html
Scroll to top button