Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

G Model

JMST-1111; No. of Pages 15 ARTICLE IN PRESS


Journal of Materials Science & Technology xxx (2017) xxx–xxx

Contents lists available at ScienceDirect

Journal of Materials Science & Technology


journal homepage: www.jmst.org

A state-of-the-art review on passivation and biofouling of Ti and its


alloys in marine environments
Shaokun Yan a , Guang-Ling Song a,c,∗ , Zhengxian Li b , Haonan Wang b , Dajiang Zheng a ,
Fuyong Cao a , Miroslava Horynova a , Matthew S. Dargusch c , Lian Zhou d
a
Center for Marine Materials Corrosion & Protection, State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Materials, Xiamen
University,422 S. Siming Rd., Xiamen 361005, China
b
Corrosion & Protection Research Lab, Northwest Institute for Non-ferrous Metal Research, Xi’an 710016, China
c
Centre for Advanced Materials Processing and Manufacturing (AMPAM), School of Mechanical and Mining Engineering, The University of Queensland,
Brisbane, Qld 4072, Australia
d
Nanjing Tech University, Nanjing 211800 China

a r t i c l e i n f o a b s t r a c t

Article history: High strength-to-weight ratio, commendable biocompatibility and excellent corrosion resistance make
Received 11 August 2017 Ti alloys widely applicable in aerospace, medical and marine industries. However, these alloys suffer
Received in revised form 4 November 2017 from serious biofouling, and may become vulnerable to corrosion attack under some extreme marine
Accepted 6 November 2017
conditions. The passivating and biofouling performance of Ti alloys can be attributed to their compact,
Available online xxx
stable and protective films. This paper comprehensively reviews the passivating and biofouling behavior,
as well as their mechanisms, for typical Ti alloys in various marine environments. This review aims to
Keywords:
help extend applications of Ti alloys in extremely harsh marine conditions.
Ti alloy
Passivity © 2017 Published by Elsevier Ltd on behalf of The editorial office of Journal of Materials Science &
Biofouling Technology.

1. Introduction hydrocarbon extraction devices in offshore petrol-chemical indus-


try, heat exchangers in desalination plants, and cooling systems
Ti alloys have been identified as superior corrosion-resistant in seawater-cooled power plants [7–9]. Many researchers [10–13]
lightweight structural materials in aerospace applications. These have compared the corrosion rates of Ti alloys with those of other
materials can also reduce the weight of components in transport metallic materials in marine environments, and found that the
applications and thus the consumption of fuel [1–3]. For example, measurement and evaluation of short-term corrosion damage of Ti
Ti alloys can be made into connecting rods, valves, wheel rim screws alloys are relatively difficult due to their high corrosion resistance.
and suspension springs for racing cars [4], and used for tail covers However, Ti has a serious biofouling problem because of its
and turbine blades in aircraft engines [2]. Because of their excellent excellent biocompatibility. In marine engineering, biofouling is an
corrosion resistance, they are even commonly used as permanent issue as critical as corrosion damage. Most investigators are partic-
orthopedic and dental implants [5,6]. ularly interested in the adherence mechanism of bacteria on Ti in
The excellent mechanical strength and chemical stability of Ti seawater and the corresponding antifouling strategies [14–17], but
alloys along with the opportunity for weight reduction make these they often ignored the effect of the surface conditions or state of Ti
materials ideal for seawater related engineering applications. For on biofouling in seawater.
example, they have been made into fasteners in marine industry, It is true that Ti and its alloys are highly corrosion resistant in
general environments. Yet in certain service conditions, failure of Ti
components does occasionally occur. For example, the bond inter-
faces of heat exchangers made of grade 3 Ti were deformed and
∗ Corresponding author at: Center for Marine Materials Corrosion & Protection, eventually destroyed by hydride in seawater [18]. Similarly, tubes
State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Materials, made of grade 2 Ti, exposed to a natural seawater environment
Xiamen University,422 S. Siming Rd., Xiamen 361005, China.
within heat exchangers, ruptured due to hydrogen assisted corro-
E-mail addresses: guangling.song@hotmail.com, glsong@xmu.edu.cn
(G.-L. Song). sion (HAC) [19]. Even worse, corrosion and biofouling may occur

https://doi.org/10.1016/j.jmst.2017.11.021
1005-0302/© 2017 Published by Elsevier Ltd on behalf of The editorial office of Journal of Materials Science & Technology.

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
2 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

concurrently. Ti per se can have a biofouling load about 0.56 g/cm2 3. Oxide films
in field [20]. Simultaneous corrosion would definitely multiply this
damage. It was reported that a large size of perforation occurred in 3.1. Native oxide film
a Ti condenser tube, which was suffering serious fouling in a sea-
water environment, as a result of the erosion-corrosion caused by The high corrosion resistance of Ti can be attributed to its surface
turbulent flow in the biofouling area [21]. oxide film. Such a corrosion resistant film on Ti is a thin oxide layer
In nature, corrosion and biofouling could be correlated, as they spontaneously formed on the surface in atmosphere or an aqueous
are both closely associated with the surface state or film of a Ti solution, or obtained through a special process, such as anodic oxi-
alloy. However, many investigators simply looked at them sepa- dation or hydrothermal treatment. From a thermodynamic point
rately. So far, corrosion/passivity and biofouling of Ti alloys have of view, Ti can rapidly react with oxygen and produce stable Ti
never been reviewed together, and the passivating and biofouling oxides. The air-formed incipient film, ca. 1 ∼ 8 nm thick, is primar-
performance of Ti alloys in emerging marine applications have not ily composed of amorphous TiO2 with a band gap energy of 3.05 eV
been comprehensively summarized. [37,43,44], which can be affected by the ambient environment to a
This review paper provides a brief description of the physical- certain extent [43,45–47]. The reaction equation can be expressed
chemical properties of Ti alloys before describing the primary as follow [37]:
mechanisms governing the growth of oxide films on Ti surfaces.
Ti + O2 = TiO2 (1)
Following that, the paper further summarizes various types of
corrosion damage in marine environments, describes microbiolog- Thickness of the oxide film exposed in the air increases with
ically influenced corrosion (MIC) and microbiologically influenced time. Chan et al. [48] reported that a chemisorbed amorphous film
corrosion inhibition (MICI) phenomena, and lists some anticorro- was formed on the transition metal surface along with oxygen
sion and antifouling strategies adopted in the field. The purpose adsorption.
of this review is to identify current research gaps and potential
breakthrough areas for Ti alloys for future marine engineering 3.2. Passivation models
applications.
To better reveal passivating mechanism, many models have
been proposed [49–52], among which the point defect model
2. Basic physical and chemical properties (PDM) based on migration of anion vacancies appears a reasonable
and widely accepted one. It has been used to interpret stationary
Excellent corrosion resistance and biocompatibility are two of passivation [53]. Many modified PDMs have also been reported. For
the most outstanding features of Ti and its alloys. Ti is more cor- example, a combination of PDM with Wagner’s theory [54] was pro-
rosion resistant than aluminum, stainless steel and brass [22,23]. posed for Ti in acid solutions [55]. Seyeux et al. [56] came up with
The crevice corrosion of grade 2 Ti in seawater is negligible at a generalized model on the basis of PDM by introducing a potential
temperatures up to 80 ◦ C [24]. Addition of Pd can increase the modification. This model can describe non-stationary oxide growth
crevice and pitting corrosion resistance of grade 7 and grade 11 in aqueous solution. “High-field point defect” model was recently
alloys, allowing them to be used in reducing acid environments proposed to further optimize PDM. In this model, a decrease in
[25]. Because of its excellent biocompatibility, Ti is one of the most electric field can mitigate the transport of charged lattice vacan-
popular implant materials, and has been widely used in clinical cies with increasing film thickness [57]. Different from PDM, a high
applications. For example, the adhesion of tissue and ingrowth field model (HFM), Mott-Cabrera model, has also been proposed
of orthopedic and dental implants can be improved if Ti and its based on migration of interstitial cations under non-stationary con-
alloys are used [5,26–28]. It has been reported that the commend- ditions [58]. The thickness of the oxide film according to HFM can
able biocompatibility of Ti implants generally comes along with be infinite if the applied voltage is high enough, which is unreason-
good in vivo bacterial adherence, which could result in failure of able. Considering the semi-conduction of charge carriers and the
transplant operations [29–32]. transportation of anions and cations, Song proposed a passivation-
The high corrosion resistance of Ti is attributed to the compact transpassivation-second passivation model for 304 stainless steel
and stable oxide film spontaneously formed on its surface in air [51], which successfully explained the increase in film thickness
or in oxygen containing environments. Thermodynamically, pure and EIS changes with increasing potential in the passive, transpas-
Ti is reactive having a standard electrode potential as negative as sive and secondly-passive regions. For Ti at high anodic potential
−1.63 VSHE [33]. Its surface is normally covered by a passive film, region, the current density peak may also be predicted by this
which is intact in most ordinary conditions and can effectively sep- model.
arate Ti from environmental medium. Once the film breaks down, Experimental data have showed that PDM is more appropriate
it can be rapidly repaired in the broken area. Rapid corrosion may for Zr, W, Ta etc. than HFM [52]. It may also be applicable for Ti,
occur once its protective film decomposes under an anaerobic envi- as the oxide film on Ti is a multi-layer on Ti, which cannot be
ronment. Ti is not completely immune to corrosion in acid media described by HFM. However, according to the newly developed
(see Fig. 1) [34], especially in reducing acidic solutions. It has been PDM, film resistance should be dominated by a porous outer layer
reported that pitting corrosion can occur in solutions containing F− [53,59]. This inference is contrary to the Ti oxide structure. There-
or Br− , and some organic solutions like formic acid (and formates) fore, more investigations may still be needed in order to establish
in certain conditions [35–38]. In fairly strong oxidizing media of a more comprehensive film model for Ti [53].
dry chlorine and fuming nitric acid, Ti can also react fiercely as its
oxide film is not stable in this case [39]. To improve fatigue resis- 3.3. Potential dependent oxide film
tance and wear resistance, ␤ phase stabilizers are sometimes added
into Ti [40]. In developing new biocompatible Ti alloys, toxicity Anodic oxidation can effectively grow a film on Ti, and has
and modulus should be carefully specified in addition to passivity become one of the most prevalent, simple and low-cost electro-
[41,42]. chemical treatments [60]. The applied anodization potential mainly
Since many physicochemical properties of a Ti alloy are deter- drops across the oxide film. The high electric field derived from the
mined by its surface passive film, it is most likely that this film also potential drop impels the migration of Ti4+ and O2− ions to form
critically determines the passivating and biofouling behavior. TiO2 and accelerates the growth of the oxide film [60]. Delplancke

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 3

Fig. 1. Potential-pH diagram for titanium-water system at 25 ◦ C (TiO2 and Ti2 O3 are in the hydrated state) [34].

et al. [61] indicated that the oxide film had an amorphous structure
when the anodic potential was lower than 20 V. However, when
the potential was higher than 45 V, a few anatase or rutile crys-
tals could be detected. The crystal transition was dependent on
experimental conditions, such as solution temperature, deposition
time and modulated potential. The crystallization of the oxide film
on Ti could be associated with the incorporation of water at high
temperature [62]. The oxide film could vary from an amorphous
structure to anatase crystal, and then to rutile crystal with increas-
ing potential. It has been reported that the thickness of the oxide
film increases linearly with applied potential [36,46,60]. Under a
high voltage, the film thickness can be more than 1 ␮m [63]. It has
also been reported that the film thickness is dependent on the grain
orientation of the Ti substrate [64].
Fig. 2. Three-layer model for an oxide film on Ti surface [43].

3.4. Composition and microstructure of surface oxide film


reduce TiO2 layer to Ti2 O3 , followed by further reducing the latter
The oxide film is often described as a complex multi-layer struc- to TiO. It has also been assumed that transitional layers of different
ture. Its composition and microstructure depend on environmental valence states, even non-stoichiometric oxides, exist in the inter-
parameters, such as temperature and pH. A native oxide film often mediate section of the film [43]. Azumi et al. [68] proposed that
becomes inhomogeneous as it grows. Müller et al. [65] proposed TiO3 (Ti4+ O2− , O2 2− ) would be thermodynamically formed near-
that a film on Ti consisted of two discrete oxide layers. Johansson surface rather than evolution of oxygen at above 3 V during anodic
et al. [66] showed that more than one oxide species on the surface oxidation:
of Ti could be detected by XPS spectroscopy. The first systematic TiO2 + OH− → TiO3 + H+ + 2e− ( > 3 V) (2)
and quantitative research on the nature of the oxide film on Ti was
conducted by Carley et al. [45], who demonstrated that Ti2+ , Ti3+ where  refers to the electric potential (vs. RHE). However, they did
and Ti4+ species were present in a growing oxide film. Ti4+ was not provide direct experimental evidence to support the generation
the dominant portion of the outermost layer, and the lower oxi- of TiO3 .
dation states Ti2+ and Ti3+ were mainly formed at the metal-oxide
interface as an inner barrier layer [67]. This viewpoint has now 3.5. Semiconducting property
been extensively accepted. A physical model for the oxide film with
three types of Ti species has been proposed; the inner two lay- The passive film on a metal is an imperfect semiconductor.
ers consisting of Ti2 O3 and TiO, respectively, and the outer layer Reducing the defect concentration in the film will lead to a lower
of TiO2 . A schematic diagram of the “three-layer model” of oxide ion transport rate, resulting in higher corrosion resistance. Many
film is presented in Fig. 2 [43]. Argon ion sputtering might partially oxide films can be treated as a semiconductor and the distribu-

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
4 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

tion of charge carriers be described by the Mott-Schottky model.


The donor density (Nd ) may be related to the stability and integrity
of the passive film. More protective TiO2 films may be obtained
through a thermal or micro-arc oxidation treatment. Nd of the pas-
sive film after the treatment (∼1015 cm−3 ) was found to be much
lower than that of the natural one (∼1019 cm−3 ) [69]. The decreased
donor density in the depletion layer of the film can be attributed
to the enhanced migration of interstitial Ti3+ ions or ionized oxy-
gen vacancies [70]. A higher density of grain boundaries produced
by grain refining could result in a dense film formed on Ti alloys,
which could decrease Nd [71]. Although Ti oxide is usually con-
sidered as an n-type semiconductor, its p-type semiconducting
properties can also be detected in some conditions. Wang et al. [72]
synthesized a Ti-defected TiO2 with p-type conductivity through
a special thermal treatment. This defected structure had a high
charge mobility. With this oxide, Ti had worse passivity. Similarly,
p-type semiconductivity could be obtained if Ti vacancies were
generated through rapid cooling [73]. The two different types of
semiconductors possess different donor carries, and have different Fig. 3. Critical temperatures of crevice corrosion for different materials in chloride
containing solutions [89].
corrosion, photocatalysis and passivity performance.

Table 1
3.6. Modification of oxide film and its properties
Crevice corrosion damage of pure Ti coupled with gasket materials in de-aerated 6%
NaCl solution [90].
Many surface treatments, such as electrophoresis [74,75], laser
Specimen Crevice Corrosion
deposition [76], plasma spraying [77], and sol-gel dipping [78,79],
have been adopted to improve the biocompatibility of Ti. Numer- 100 ◦ C 80 ◦ C 70 ◦ C
ous studies have been carried out on the properties of TiO2 film at Ti/Ti 䊉  
different potentials [80–83]. Potential is an important factor that Ti/Neopren rubber/Ti 䊉䊉  
can significantly affect the composition, microstructure and prop- Ti/Asbestos/Ti 䊉䊉 䊉 
erties of a passive film on pure Ti and its alloys. Souza et al. [81] Ti/Teflon/Ti 䊉䊉 䊉䊉 
Ti/Dimethacrylate/Ti 䊉䊉䊉䊉 䊉䊉䊉 䊉䊉
prepared a very thick Ti oxide film at a modulated anodic poten-
tial. Wake et al. [9] applied a fixed potential around 1.0 V versus : Crevice corrosion did not occur.
䊉: Crevice corrosion occur.
Ag/AgCl to prevent Ti from biofouling in natural seawater. Such an
anodic polarization exhibited a commendable antifouling effect on
Ti, probably due to formation of a thin amorphous TiO2 film and in chloride containing solutions. A comparison in crevice corrosion
generation of oxygen radicals. resistance between Ti and several other metals in chloride contain-
ing solutions at different temperatures is depicted in Fig. 3 [89]. In
4. Dissolution in halide containing solutions solutions containing 1 wt.% Cl− , crevice corrosion occurs on pure Ti
only at temperatures higher than 100 ◦ C, much higher than that on
The dissolution of a metal is to a great extent controlled by its stainless steels, while some Ti alloys like Ti-0.15Pd will not suffer
surface film, and localized corrosion usually results from break- crevice corrosion even above such a temperature. Satoh et al. [90]
down of the film. Halides are very detrimental to a passivated metal, immersed Ti specimens with crevices, coupling with gasket mate-
as they can accelerate the breakdown of passive films. Therefore, rials in deaerated 6% NaCl solution for 720 h at three temperatures.
the dissolution of Ti and its alloys can be influenced by halides in Their results (see Table 1) show that the Ti/dimethacrylate/Ti was
solutions. The high chloride content is one of the main character- most susceptible to crevice corrosion, and four of the parallel trial
istics of the marine environment. Apart from chloride, seawater specimens were all subjected to corrosion at 100 ◦ C.
also contains some other halides. Although their concentrations are
considerably lower than that of chloride, they can also influence the 4.2. Fluoride
corrosion of Ti. The corrosion damages of Ti and its alloys exhbit dif-
ferent forms or types, depending on the metallurgical factors and Ti is more susceptible to corrosion in acidic fluoride solutions.
the environment conditions. Even with a trace amount of fluorine, remarkable damage will
appear on Ti. This can be attributed to a complex interaction
4.1. Chloride between F− and TiO2 [91]. The strong reaction of Ti oxide film with
F− in strong acid media can be expressed as below:
In the initial stage, the pitting corrosion of Ti induced by halide TiO2 + 6F− + 4H+ → TiF6 2− + 2H2 O (3)
ions can be explained by the specific adsorption of aggressive
anions instead of the point defect model. The oxide film becomes The protective film on Ti-6Al–4V can be destroyed by fluoride
thinner after adsorption of these anions. Eventually the substrate ions due to formation of Na2 TiF6 if NaF concentration reaches 0.1%
metal is exposed locally and thus rapidly dissolved there [84,85]. [92]. At the incipient stage, the oxide film is uniformly dissolved.
The breakdown potential (Eb ) of the passive film on a Ti alloy is Thereupon, bare Ti vigorously reacts with F− . The reaction can be
usually much higher than those of other metals [86]. For exam- expressed as follow [36]:
ple, Eb of Ti was reported to be ca. 12 VSHE in 0.1 M NaCl, much
2Ti + 12F− + 6H+ → 2TiF6 3− + 3H2 ↑ (4)
higher than that of stainless steel [87]. It is generally accepted that
the oxide films on Ti and its alloys are extraordinarily stable in The film deterioration can be affected by environmental pH. The
environments containing invasive chlorine ions [88]. Crevice cor- lower pH, the more HF compounds will be formed, and the corro-
rosion is also more difficult to occur on Ti than on stainless steels sion damage will be more serious. The critical HF concentration of

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 5

Table 2
Pitting and corrosion potentials of Ti and other alloys [108].

Sample Temperature and flow rate Pitting corrosion (SCE)/V Corrosion potential (SCE)/V

Aluminium 25 ◦ C, 0 m/s −0.76 −1.2


Aluminium brass 25 ◦ C, 15 m/s −0.1 −0.26
70/30 CuNi 25 ◦ C, 15 m/s −0.08 −0.27
304 stainless steel 25 ◦ C, 15 m/s +0.3 −0.5
316 stainless steel 25 ◦ C, 15 m/s +0.5 −0.4
Ti 25 ◦ C, 15 m/s +8 ∼ +10 −0.2∼ +0.1

30 ppm was identified when commercially pure Ti was immersed the passivation of Ti mainly depends on pH, not Cl− concentration
in solutions containing only HF [93]. The dependence of critical [104]. An inhibiting effect of SO4 2− in acid and neutral Cl− solu-
F− concentration on pH is linear at pH = 0∼2.5 and pH = 3.0∼5.0 tion at different temperatures was studied by Liu et al. [105,106].
regions, respectively [94]. Brossia et al.[95] conducted a potentio- Such results showed that the passivity of Ti in 1 M NaCl solution did
static test (0 V vs SCE) for Ti grade 7 alloy at 95 ◦ C, and found that not break down even at a high temperature when 0.5 M SO4 2− was
the critical F− concentration to accelerate dissolution of the passive added into the solution, and both the breakdown and repassivation
film on Ti was ca. 0.5 m mol/L in a deaerated 1 mol/L NaCl solution. potentials were enhanced with increasing SO4 2− concentration.
The performance of Ti and its alloys in seawater with coexistence
4.3. Bromide of Cl− and SO4 2− is not clear. So far, there is no report on corrosion
or de-passivation damage caused by F− in seawater on Ti and its
Br− is more prone to pitting corrosion on Ti compared to other alloys. That is probably due to the competitive adsorption of SO4 2−
halide ions, and the pitting corrosion could occur on Ti at ∼1.4 V or other oxygen-containing ions against F− in seawater. However,
vs Ag/AgCl in neutral solutions [96,97]. Ittah et al. [97] reported some electrochemical results have indicated that grade 2 Ti has a
that the pitting potential of Ti grade 2 was strongly dependent on corrosion rate of 0.1 mpy in seawater containing 0.5% H2 SO4 [107].
the concentration of Br− ions, and a linear relationship between
pitting corrosion rate and NH4 Br concentration below 4 mol/L had 5. Corrosion in marine environments
been obtained. However, the monotonic concentration dependent
relationship could not be extended to a concentrated solution, as 5.1. Crevice corrosion
Ti became insensitive to pitting due to the absence of water in the
medium when the concentration reached the saturation point. The Ti and its alloys are not only resistant to chlorides, but also to
anomalous pitting sensitivity of Ti may be related to the higher many other species in natural seawater. Pitting potential of Ti in
polarizability of Br− ions which donate electrons more readily flowing seawater is much higher than other alloys at 25 ◦ C. Their
than chloride ions and increase the electrical conductivity of pas- pitting and corrosion potentials are listed in Table 2 [108], which
sive films [98]. Huo et al. [99] claimed that impurities (FeTi and shows that pitting phenomenon can rarely be found on Ti in a
FeTiO3 ) existing on the surface of Ti could act as active points where marine environment.
anodic current density was high, pitting nucleation energy was low, Ti and its alloys are relatively susceptible to crevice corrosion
and adsorbed Br− concentration increased with increasing voltage. in anaerobic seawater at high temperatures. The crevice corrosion
When the number of Br− nuclei exceeded a critical surface con- resistance of Ti and other metals exposed to static seawater are
centration, pitting corrosion occurred. The rate-determining step compared in Table 3 [109]. A conclusion can be drawn that Ti and
of the reaction could be written as follow: Hastalloy C have better crevice corrosion resistance than the other
TiO2 + 4Br− → TiBr4 + 2O2− (5) metals. He et al. [62] investigated the influence of temperature on
crevice current and potential using a galvanic coupling technique.
where TiBr4 could be easily hydrolyzed and TiO2 deposited again: The results showed that there was a temperature threshold around
TiBr4 + 2H2 O → TiO2 + 4H+ + 4Br− (6) 70 ◦ C to initiate crevice corrosion for Ti in hot halide or sulfate con-
taining solutions. In seawater, crevice corrosion would not occur
The generated H+ could accelerate the dissolution of TiO2 : at temperatures below 80 ◦ C [88]. The crevice corrosion of Ti is an
+
TiO2 + 2H → TiO 2+
+ H2 O (7) autocatalysis process, similar to other metals. The reactions within
the crevice can be expressed as follows [62]:
The surface binding energy for halide ions has been obtained
through Langmuir isotherm, and the results suggest that Br− ions Ti4+ + 2H2 O ↔ [Ti(OH)2 ]2+ + 2H+ (8)
are selectively adsorbed [84]. Ti tetrabromide or oxybromide can
Ti4+ + 4H2 O ↔ Ti(OH)4 + 4H+ (9)
be formed at the oxide/solution interface as intermediate products.
In fact, the Ti tetrahalide or oxyhalide generation model has been Breakdown and/or recrystallization of the oxide film may occur
proposed in Ti pitting reaction systems [96], and the validity of the at a high temperature. This may lead to formation of open path-
point defect model has been further examined by Sazou et al. [100]. ways in the grain boundaries of the film, resulting in dissolution
of substrate metal. The crevice corrosion of Ti at low and high
4.4. Sulfate temperatures can be schematically illustrated as shown in Fig. 4
[62].
SO4 2− is an aggressive anion to Ti, and should not be ignored due The redox process inside the occluded crevice can be facilitated
to its great amount in seawater. It has been reported that the gen- by cation hydrolysis and local acidification. A similar conclusion
eral corrosion of Ti can be accelerated by increasing concentration concerning acidification in the crevice has also been drawn by
of SO4 2− (0.001 ∼ 10 mM) in oxalic acid [101]. Localized corrosion Rajendran et al. [110]. In addition, HCl is probably generated due to
has been found on Ti alloys after long-term immersion in sulfate the existence of MgCl2 in seawater through the following reactions
acid [102]. Oxygen-containing anions usually act as inhibitors for [111]:
pitting corrosion of metals in halide solutions due to their preferen-
tial adsorption on the surface oxide films [103]. In H2 SO4 solutions, MgCl2 ·2H2 O ↔ MgOHCl + HCl + H2 O (10)

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
6 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

Table 3
Crevice corrosion resistance of Ti and other metals in static seawater [109].

Crevice corrosion resistance Crevice induced serious pitting

Excellent Good Moderate Inferior

Ti 90 ∼ 10 CuNi1.5Fe Cast iron Incoloy 825 AISI 316, 304, 400 series
Hastalloy C 70 ∼ 30 CuNi0.5Fe Carbon steel NiCu alloy stainless steel
Bronze Cu
Brass

alloys exposed to the deep water may have different surface states
compared to that in the shallow sea or atmospheric marine envi-
ronments. It has been reported that in deep seawater environments
Ti corrosion might be facilitated because of high pressure [116].
Moreover, the crevice corrosion of Ti can be significantly affected by
temperature. Pang et al. [24] found that grade 2 Ti started to exhibit
crevice corrosion features at 80 ◦ C and the corrosion penetration
depth reached about 90 ␮m at 200 ◦ C, but for Ti-6Al–4 V no crevice
corrosion was observed below 200 ◦ C. They further investigated the
effect of CO2 on crevice corrosion and found that corrosion penetra-
tion depths evidently increased at temperatures lower than 200 ◦ C
compared to their former work, especially for grade 5 Ti in the pres-
ence of CO2 [117]. Crevice corrosion does not always occur when
temperature increases further. For example, no crevice corrosion
has been found on grade 2 Ti and Ti-6Al–4 V samples at 232 ◦ C, but
has been detected at lower temperatures [88]. Ti-Pd, Ti-Ni, Ti-Co
and Ti-Mo are four typical enhanced Ti alloy systems showing high
crevice corrosion resistance in marine environments. The crevice
corrosion results of these Ti alloys at high temperatures in arti-
ficial seawater are listed in Table 5 [118]. Ti-Co and Ti-Mo alloys
with increased alloying element additions do not exhibit crevice
corrosion.
Micro-arc oxidation (MAO) treatment and anodic polarization
can also increase the crevice corrosion resistance in acid brine solu-
tions. The influence of these factors on Ti alloys under different
acidity levels and temperatures is depicted in Fig. 5 [7]. Ti alloys
could react with hydrogen atoms and reach a hydrogen-saturated
state in high pH and high temperature media. This reaction may
occur when Ti alloys are used in concrete as rebar in some practical
Fig. 4. Crevice corrosion of grade 2 Ti at (a) low temperatures, and (b) temperatures service environments.
greater than 65 ◦ C [62].

Table 4
5.2. Galvanic corrosion
Corrosion rates of pure Ti and Ti-6Al–4 V at different depths in seawater
[39,113–115]. The driving force for galvanic corrosion primarily depends on
Sample depth/m corrosion rate (mm/y) Reference
the open-circuit potential difference between coupled metals. Ti
is nobler than other engineering metals in seawater. The galvanic
pure shallow 0.8 × 10ˆ−6 39
Ti
series of some metals in ambient seawater is shown in Fig. 6
720 ∼ 2070 <0.00025 115
1300 ∼ 1370 <0.00025 115 [23]. The open-circuit potential of Ti in natural seawater can shift
1.5 ∼ 2070 0.0 115 towards a quasi-steady-state value of ∼50 mV vs SCE, as a result of
1720 0.00004 113 an oxide film formed and thickened on Ti [119,120]. The galvanic
TC4 1.5 ∼ 2070 <0.00025 115 potentials of Ti coupled with a single metal or multiple metals in
1720 0.8 × 10ˆ−6 113
1720 ≤0.001 114
natural seawater are schematically illustrated in Fig. 7. The gal-
vanic potentials shift to the anodic electrode potential eventually
[121]. Therefore, the corrosion rate of a less noble metal will be
MgCl2 ·H2 O ↔ MgOHCl + HCl (11) accelerated when coupled with Ti. Galvanic corrosion rate is usually
governed by anode materials. Du et al. [122] showed that copper
MgOHCl ↔ MgO + HCl (12)
acted as anode in contact with Ti and was corroded after long-term
The generation of HCl can reduce the pH value in the occluded immersion, and its corrosion product hydroxyl copper eventually
crevice and accelerate the electrochemical reactions. peeled or dissolved, resulting in cracking and pitting damage. The
Ti may be general corrosion resistant in deep seawater environ- variation trend of the galvanic corrosion potential was similar to
ment. For example, deep sea well-heads made of Ti stand a good that of a single copper coupon. Din et al. [121] believed that O2
chance of surviving in anaerobic and high temperature (∼ 140 ◦ C) dissolved in seawater was a depolarizer which consumed excess
conditions [112]. The corrosion rates of pure Ti and Ti-6Al–4 V at electrons that could not be used in H2 evolution reaction in a gal-
different depths of seawater are listed in Table 4 [39,113–115]. They vanic couple. The oxygen reduction reaction can be expressed as:
are fairly low and not significantly different at different depths
for both pure Ti and Ti-6Al–4V. However, the oxide films on Ti O2 + 2H2 O + 4e− = 4OH− (13)

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 7

Fig. 5. Effects of temperature, acidity, alloying and surface treatment on the susceptibility of Ti alloys to crevice corrosion in chloride solutions [7].

Hydrogen can reside in Ti in two forms, either as free atoms or retarding effect of the interaction has also been observed in some
as hydride TiH2 . It can only desorb at a relatively positive potential. cases. Ding et al. [129] reported that the volume loss of TC11 in
Sacrificial anode protection is based on the well-known gal- artificial seawater was always lower than that in distilled water
vanic effect. In this couple system, surface alkalization can occur on under fretting wear testing conditions. Silicon containing natural
the cathode surface due to cathodic hydrogen evolution or oxygen seawater might form a lubricant film to reduce friction on the Ti
reduction. An increase in pH value can facilitate the conversion of surface. In a high cycle fatigue regime, seawater could improve the
HCO3 − to CO3 2− followed by deposition of CaCO3 , which has often fretting fatigue life, because more micro-cracks were generated on
been observed during cathodic protection in seawater [123]: the surface of Ti-6Al–4 V, and thus more debris could be formed as
a shield retarding the crack growth rate [128].
Ca(HCO3 )2 + 2OH− = CaCO3 +2H2 O + CO3 2− (14)

El-Dahshan et al. [124] drew a similar conclusion on the gal- 5.4. Hydrogen absorption
vanic corrosion behavior of a Ti/Al brass couple. In comparison with
the galvanic corrosion in synthetic seawater without HCO3 − , the Hydrogen embrittlement is one of the common failure modes
synthetic seawater had a larger corrosion current, indicating that triggered by hydrogen storage in metal, especially when metal is
the CaCO3 film acted as a barrier on Ti surface. Wang et al. [125] exposed in gaseous hydrogen atmosphere [130]. Hydrogen absorp-
studied galvanic corrosion of a multiphase system in which TA2 is tion can however be retarded by a highly impermeable surface
the noblest metal in seawater. The total galvanic current from the oxide film. This has been confirmed by Yen et al. [131] on a ther-
anode with the most negative open-circuit potential was equal to mally grown dense Ti oxide film. It has been reported that the
the sum of the currents to the two cathodes. formation of a thin hydride layer on the surface cannot signifi-
cantly influence the mechanical properties of Ti, but the hydride
5.3. Fretting fatigue corrosion layer can impede the process of hydrogen penetrating into Ti [132].
The amount of formed hydride is usually limited even at a polariza-
Ti is susceptible to fretting wear damage, a surface failure result- tion potential more negative than the threshold of −0.75 vs SCE at
ing from small-amplitude vibration of two solids in contact. If temperatures below 80 ◦ C [8]. An elevated temperature can accel-
corrosive medium is present in a fretting wear position, local- erate the anodic dissolution of metal and increase hydrogen activity
ized corrosion will occur after the passive film is damaged there and diffusion rate [133]. Tsai et al. [134] found that hydride precip-
[126–128]. It is generally believed that the interaction of fretting itation could occur on pure Ti surface at 50 ◦ C and −1.2 V vs SCE
wear and corrosion can facilitate the damage process. However, a in artificial seawater (pH 8.2). In a marine environment, hydro-

Table 5
Crevice corrosion damage of Ti and its alloys at high temperatures in artificial seawater [118].

Alloys Temperature/◦ C Time/h Corrosion damage degree of Teflon gasket

Ti 200 96 seriously
Ti-0.2Pd 300 500 none
Ti-Ni 200 ∼ 288 504 none
(0.1% ∼ 5%)
Ti-Co 200 96 seriously
(0.1% ∼ 0.2%)
Ti-Co 232 120 none
(0.4% ∼ 5%)
Ti-1Mo 200 96 seriously
Ti-Mo (3% ∼ 6%) 200 99 none

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
8 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

Fig. 6. Galvanic series of some metals in ambient seawater [23].

Fig. 7. Galvanic potentials of Ti coupled with single or multiple metals exposed in natural seawater [121]. (*Arrowheads refer to initial positions of the stable galvanic
potential.)

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 9

gen absorption is usually aggravated by cathodic protection. Brittle TiO2 film on Ti alloys [41,155]. Ti-22Nb-6Zr has a wide passive
hydride induced failure of Ti tubes coupled with sacrificial anodes region [156], in which Zr is a ␤ phase stabilizer similar to Mo. ␤
has been reported by Shalaby [135]. phase stabilizing elements can stabilize passive films through gen-
Tensile load and residual stress can also accelerate the aggra- erating high valent oxides of Ti alloys. A compact oxide film could
vation of hydrogen pickup and the occurrence of embrittlement be formed when 15 wt% Ta is added to a Ti based alloy, and the
[133]. Hydrogen induced cracking (HIC) is a potential threat to Ti passive film will become more resistant with increasing Ta content
and its alloys, in which cracks can preferentially propagate through [157]. Meanwhile, a continuous single phase of Ta could be formed
hydrides [136]. Nevertheless, in real seawater HIC has rarely been under the film [158–160]. A similar stabilizing effect on Ti passive
reported even under elevated temperatures because of the excel- film has also been reported on Nb [161–164].
lent protectiveness of the surface films on Ti alloys [117]. However,
due to the interaction between galvanic corrosion and crevice cor-
7. Biofouling
rosion [19], the hydrogen assisted corrosion (HAC) of Ti may be
initiated at temperatures 35 ∼ 40 ◦ C in a joining area, 30 ∼ 35 ◦ C
Surface films can influence the adhesion of external species,
lower than that in other areas.
including bacteria and fungi. Biofouling in general can be described
as undesired adhesion and growth of invisible aquatic bacteria and
5.5. Other possible corrosion damage fungi, or of a community of visible plants and animals on a material
surface [165]. It can be classified into microfouling and macrofoul-
In marine engineering applications, some other types of cor- ing [166].
rosion could also occur on Ti and its alloys, such as stray current Investigators believe that biofouling on a metal surface always
induced corrosion (the stray current may come from a cathodic leads to an increase in corrosion rate [167–169]. However, this is
protection system), cavitation corrosion (particularly when water not always true. Same species may show corrosive or protective
flow speed is too high or Ti alloy components are moving too fast in behavior under different fouling conditions. The former was usu-
water), corrosion fatigue (resulting from wave impact etc.), erosion ally known as microbiologically influenced corrosion (MIC) [170],
(when solid particles from sea mud are involved), and sea creature while the latter was defined as microbiologically influenced corro-
induced corrosion. As Ti is yet to be widely used in marine environ- sion inhibition (MICI) [171].
ments, there are only a few publications in these areas. For example,
to protect platinized Ti electrodes from attack by stray current,
7.1. Microbiologically influenced corrosion (MIC)
Kocak et al. [137] optimized the design parameters for a seawater
ground return system. Mochizuki et al. [138] studied the cavitation
MIC can endanger the reliability and security of engineering
erosion of pure Ti and its alloys in seawater, and found that the ero-
structures, which can be observed in various environments, such as
sion resistance could increase with increasing hardness. The fatigue
soil, fresh water and seawater. Roughly 20% of the corrosion dam-
crack growth rate of pure Ti exposed in seawater have been found to
ages in practice are caused by MIC [172]. Liu et al. [173] found that
be different from that in standard atmosphere environments [139].
sulphate-reducing bacteria (SRB), one of the most common cor-
Shulov et al. [140] analyzed the dent-shaped craters on the surface
rosive bacteria, accelerated the corrosion of a sacrificial anode. It
of some Ti alloys, and found that pitting corrosion was more likely
has been reported that almost all ordinary engineering materials,
to take place under cyclic loading. They concluded that the corro-
except Ti and its alloys, are prone to SRB induced MIC in anaerobic
sion was probably caused by accumulation of Cl− or a significant
environments [174].
change in composition near the craters during the fatigue tests.
Colonization of inhomogeneous surfaces and metabolic prod-
ucts of organisms are the significant factors that can influence MIC.
6. Effect of alloying on passivation The former directly influences the local surface microenvironment,
accelerating the onset of localized corrosion [175]. The latter acts
Passivation behavior of Ti can be influenced by alloying ele- as an electro-catalyst facilitating the redox kinetics on the surface.
ments. For example, if noble platinum-group metals (PGMs) are Extensive MIC studies have been focused on prokaryotic bacteria,
added into Ti, they may be enriched on the surface after Ti atoms such as sulfate-reducing bacteria (SRB), which makes use of organ-
are dissolved, and thus impel the passivation of Ti [141,142]. A ics adsorbed on metal surfaces as a source of carbon for growth
small concentration of Pd in Ti can facilitate the spontaneous pas- and reproduction, transforming SO4 2− into S2− or H2 S, and finally
sivation of Ti, as it alters the kinetics of cathodic reaction on Ti forming metal sulphides in anaerobic or low oxygen containing
surface in a reducing environment, resulting in lower passive cur- environments [174]. The pitting corrosion of Ti influenced by SRB
rent densities and improved passivity [95,143,144]. The beneficial in seawater has been found and a probable mechanism has been
effect of other PGMs can also be explained according to this mech- proposed, according to which micro-pits are generated due to the
anism [145,146]. For example, corrosion resistance of Ti could be presence of TiS2 [176,177].
improved up to 250 ◦ C due to addition of Ru or Pd without impair- The influence of fungi on corrosion has rarely been studied in
ing the alloy’s mechanical properties [8,147–149]. The catalysis comparison with the effect of bacteria. Recently, Zhang et al. [175]
of PGMs on hydrogen reduction reaction can stimulate hydrogen investigated the initial corrosion behavior of 304 stainless steel and
evolution, and consequently increase the exchange current density Ti exposed to P. variotii, A. niger and their mixtures in a high humid-
and/or reduce the cathodic Tafel slope [145,146,150]. Ru can even ity atmospheric environment for 28 days and 60 days through
be re-precipitated onto Ti surface film [141]. This further enhances Scanning Kelvin Probe (SKP) and stereomicroscopy observation.
its alloying effect on Ti passivation. They found that the presence of fungi decreased the corrosion
The insoluble molybdenum chlorides can stabilize the amor- resistance of both 304 stainless steel and Ti. The latter had higher
phous structure of titania in Ti-Mo alloys, and widen the passive resistance to corrosion than the former due to the poorer adherence
region of Ti alloys [151–153]. The analogous ennoblement effect of fungi on its surface.
of molybdenum through inhibiting the anodic reaction has been There are very few reports on the corrosion of Ti or its alloys
reported by Rajendran et al. [110] and Newman [154]. High valent induced by microorganisms [16]. Ti and its alloys in general appear
oxides, such as ZrO2 , Ta2 O5 , Nb2 O5 , are usually protective. They can to be able to remain ennobled under anaerobic conditions where
reinforce the compactness and improve the stability of the native the aggressive SRB species can easily survive. It is unclear whether

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
10 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

Ti and its alloys can still survive SRB attack in some extremely harsh electrical pulses cannot significantly affect barnacle cyprid settle-
environments. In theory, the surface oxide film on Ti can eventually ment [186].
be dissolved in contact with a strong acidic solution. Thereby, if a Calcification may provide a suitable settlement condition for
micro locally acidified environment is generated by SRB on Ti, the organisms in coral reef environments. The cathodic process
possibility of rapid MIC damage of Ti cannot be excluded. involved in the calcification process can be oxygen reduction (see
reaction (13)) and hydrogen evolution [123]:
7.2. Microbiologically induced corrosion inhibition (MICI)
2H2 O + 2e = H2 + 2OH− (15)
Corrosion inhibition due to formation of biofilms has been
After alkalization by these cathodic reactions, the subsequent
detected on some materials exposed in invasive environments in
calcification reaction will be the process described in Equation (14).
the presence of bacteria. There have been many investigations into
The cathodic reaction enhances pH value on the metal surface and
the mitigating behavior instead of the facilitating effect of prevail-
facilitates the formation of a carbonate layer for colonization of
ing SRB species on corrosion [178]. For example, Liu et al. [173]
the reef-building organisms. Vedaprakash et al. [20] compared the
speculated that the contact between a metal and corrosive media
toxicity of some metallic panels and built a toxicity tolerance base-
could be impeded by the formation of a SRB biofilm and deposition
line for primary hard-shelled biofouling organisms. They found that
of corrosion products on the metal surface, and thereby the transfer
the fouling of stainless steel and Ti was much more severe than
of electrons and invasive ions between the substrate and corrosive
the other samples, and Ti had approximately double mass gain of
media became more difficult. A similar study on aluminum alloys
the fouling load in relatively deep seawater compared with that in
[171] also showed that Shewanella could prevent pitting corrosion
shallow water.
of Al 2024.
On Ti and its alloys, the biofilm can also act as a protec-
tive shield to avoid direct contact between the substrate and the 7.4. Passivation and biofouling
bulk environmental media. Thus, the corrosion resistance should
increase accordingly. It has been widely accepted that MICI is prin- The above studies imply that the surface oxide film on Ti may
cipally responsible for the decreasing rate of oxygen reduction be intact during the biofouling process, and macro-organisms tend
triggered by microbial respiration on metal surfaces [171,179]. to stay on the intact oxide film. Although this point has not been
Ameer et al. [180] measured the inhibition efficiency (IE%) of bio- systematically verified by direct evidence so far, the fact that metal
logically active substances extracted from marine red algal on surface modification by a passive layer can significantly influence
Ti-6Al–4 V, and found that the active substance had the highest organism adherence has been well known. There is clear inter-
inhibition efficiency, which could be attributed to the synergetic action between bacteria, medium and materials, including the
effect of antioxidative ingredients in J3 and scavenging free radicals evolution of semiconducting properties of passive layers on met-
present on Ti-6Al–4 V surface. als, in marine environments [187]. For example, ferrous metal could
If the surface passive film of metal is perfect, the MICI effect even be dissolved and combined with extracellular polymeric sub-
cannot be significant. The reported enhancement in IE may be asso- stances (EPS), comprising proteins, polysaccharides, nucleic acids
ciated with the repaired surface film in some defective points by the and lipids, secreted by adsorbed cells, and formed Fe3+ (EPS) com-
biofilm or biologically active species. pounds [188]. Since metallic ions may be involved in metabolism of
adhered organisms, the electrochemical reactions in theory should
7.3. Macrofouling on Ti and its alloys be able to influence the activity of a biofilm on the metal surface.
Therefore, an overdose of metallic ions is definitely detrimental
Only a trace amount of Ti can be released to biotic environ- to the growth of a biofouling film, and an inert surface should be
ments due to the presence of a compact protective surface film. vulnerable to biofouling.
This low-toxic or non-toxic characteristic makes Ti highly suscep- For Ti, its passivity is a result of a protective film formed on the
tible to macrofouling. Although Ti and its alloys are immune to MIC surface, which effectively separates the substrate from its environ-
in most situations, their corrosion and mechanical performance can mental medium or makes the substrate inert. Thus, metallic cations
be influenced by macrofouling. Fouling organisms mainly include cannot be easily released from Ti to react with organic molecules in
barnacles, mussels, oysters, bryozoans and algae, which can reduce biofilms. On the other hand, many species from the biofilm can still
the speed of vessels and the efficiency of fuel, and increase the change the microenvironment at the interface between the biofilm
maintenance cost if they are attached to hulls [3,165,166]. The and passive film, resulting local acidification and depassivation of
heat transmission and flow in Ti heat exchanger tubes can even Ti, which may even accelerate cathodic reaction [177,189].
be dramatically diminished under fouling conditions [181,182]. It Many factors can influence the biofouling behavior. For exam-
was found that the flow in Ti tubes with 24.4 mm internal diameter ple, oxygen was regarded as an electron acceptor producing H2 O2
was reduced by over 40% due to biofouling after one-year service in for cathodic reduction, resulting in surface acidification [190–193].
natural seawater [183]. The general corrosion rate of Ti exposed in Sulphate can also be an acceptor for SRB. In that case, sulfide can
natural seawater has been reported to be higher than that in ster- be produced on a passive film surface, which can eventually lead to
ile seawater due to the effects of macrofouling (see Table 6) [184]. localized damage [177]. Passive film thinning has also been found in
It has been reported that grade 2 Ti exposed to a seawater envi- the presence of SRB biofilms. Moreover, surface charge can also be a
ronment is subjected to negligible corrosion, but has substantial significant factor determining the adhesive performance of marine
biomass gain at a rate of 0.4527 kg m−2 y−1 [184]. Thus, macrofoul- bio-organisms. If the surface charge is present on a biofilm, it can
ing is a main threat to the successful application of Ti in marine influence the corrosion resistance of the metal through repulsing
environments. corrosive ions [188,194].
Most of the studies on the macrofouling of Ti and its alloys have
been conducted by means of field tests, which normally take several 8. Antifouling strategy
months or even years [184,185]. It appears that Ti and its alloys are
nearly corrosion free while undergoing biofouling. When a cathodic Numerous studies have been conducted to explore more
potential −10 V vs SCE is applied, the average percentage of settled economic, environment-friendly and effective techniques to
barnacles will be noticeably reduced, whereas smaller amplitude prevent Ti from biofouling. The fouling substances can vary

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 11

Table 6
Electrochemical parameters of Ti immersed in seawater [184].

Medium Sample No. Duration(months) Ecorr , mV Icorr uA/cm2 ba, mV/dec bc, mV/dec Corrosin rate, mmpy

NSW 1 3 −175 0.22 – – 0.0037


2 6 −71 0.26 66 48 0.0044
SSW 1 3 −145 0.027 13 19 0.0005
2 6 −70 0.065 19 17 0.0011

NSW: natural seawater.


SSW: sterile seawater.

Fig. 8. Marine biofouling organisms and their morphology and size evolution with time [196].

from attached low-grade lives to subsequently adhered high- Although these reagents mentioned above are effective in
grade lives during the whole process [195]. The evolution of governing the biofouling of Ti in seawater, their durability and
marine biofouling organisms and their morphology and size long-term environmental impact need further assessment.
with time are depicted in Fig. 8 [196]. The formed microbial
biofilms are less vulnerable to antibacterial agents [197,198]. 8.2. Electrochemical protection
An effective control of the initial microbial adhesion, followed
by measures to retard the settlement and metamorphosis of There is a tendency to replace paints containing heavy metals
hard-shelled organisms, are the keys to successful antifouling with environment-friendly antifouling technology. Electrochem-
[199]. ical inactivation is an effective method in controlling bacterial
cell accumulation and marine fouling [206–208]. Matsunaga et al.
8.1. Toxic additives [206,209] indicated that the electrochemical passivation was a
non-toxic process, as it was a direct electron transfer between
Many researchers are still trying to improve the antifouling per- microbial cells and electrode. However, many electrochemical
formance through incorporating toxic constituents, such as organic approaches are not suitable for practical applications due to rapid
tin and heavy metal ions into coatings or the material surface physical degradation of the electrodes under the applied poten-
directly. Tributyltin (TBT) was widely used for antifouling in the last tials. Cathodic polarization can influence both anticorrosion and
century due to its eximious performance of killing settled fouling antifouling of Ti. The cathodic oxygen reduction and the electro-
organisms and the so-called self-polishing behavior. However, TBT static effect caused by surface charge may be responsible for the
has been prohibited in many countries for its residual toxic effect electrochemical inhibition on bacteria [210]. A micro-porous struc-
on non-targeted organisms [3]. Copper is abundantly applied in ture is usually made on Ti under a high anodic potential to obtain
antifouling as biocide for its toxicity resulting from Cu2 O released a hydrophobic surface for anti-bacteria [15,17,211]. Wake et al. [9]
into the ambient media, which can hinder the process of cell divi- used a stable electrode material to make an electrochemical sys-
sion [200–203]. Originally, copper was directly used on the bottom tem for antifouling in cooling lines. A Ti electrode was used in the
of ships to impede biofouling [166,204]. Later on, antifouling paints electrochemical antifouling system with applied potentials around
containing copper salts were popularly applied in marine engineer- 1.0 V vs Ag/AgCl. Field test results showed that the antifouling was
ing. Vishwakarma et al. [200] prepared a few copper-nickel bilayer effective and the antifouling rate was approximately 80%.
and multilayer thin films on Ti through a pulsed laser deposition There are no needs to handle the spalling of paints yearly if elec-
technique. The Ni layers served as barriers to delay the leaching of trochemical inactivation methods are used. However, the on-going
Cu+ ions and enhance the persistence of the antibacterial layer. The energy consumption for the electrochemical inactivation is costly.
large surface area of the (Cu/Ni)10 multilayer had a more favorable
antibacterial property compared to the bilayer system. An inhibi- 8.3. Hydrophobic surface modification
tion effect of 3% Ag+ doped titania on MIC and corrosion in seawater
has been confirmed. Sterilizing rate of the doped TiO2 reached Hydrophobicity was found on the leaves of lotus and taro and
97.1%, higher than that of the undoped TiO2 after sterilizing for the wings of cicada in nature [212–214]. A (super)hydrophobic
24 h [205]. Ti surface can be obtained by different treatments. For example,

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
12 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

the water contact angle decreased with increasing voltage used Ti, including its thickness, composition, microstructure and semi-
in surface conversion treatment and ultraviolet (UV) light illu- conducting characteristics. However, the passivation property of Ti
mination, and annealing treatment can also be a crucial factor and its alloys in marine environment has not been systematically
affecting the hydrophobicity [83]. Hydrophobicity will be sub- studied. Ti substrate can react fiercely, particularly inside crevices
stantially enhanced if surface free energy decreases and surface in anaerobic seawater at high temperatures, once the oxide film is
roughness increases [215–217]. A superhydrophobic Ti surface damaged, but relevant reports on corrosion damage of Ti and its
has preferable antibacterial performance and corrosion resistance. alloys are very limited.
Such a surface can limit the direct contact between Ti and media, as The effect of halide and sulfate ions in seawater, even at a low
air is likely to be trapped in the nanopores in the surface oxide film concentration, on Ti passivity should be carefully investigated. Pit-
[218,219]. Superhydrophobicity can be attributed to the hierarchi- ting corrosion of Ti may occur in neutral bromide solutions probably
cal micro- and nano- structures of the surface [5,220]. This kind of due to the selective adsorption of halide ions on Ti surface, and the
surface is prevalent in medical implants to prevent biofilm-induced corrosion products are mainly Ti tetrahalide and oxyhalide. The
infections. To attain antibacterial properties, some investigators detrimental effect of halide ions may be restrained by preferential
modified the surface of Ti inspired by bionics. For example, Bhadra adsorption of sulfate ions.
et al. [32] cultured S. aureus and P. aeruginosa, two conventional Alloying elements can affect the corrosion and passive perfor-
pathogenic bacteria, on Ti surfaces with nano-patterns that are mance of Ti and its alloys. The passivity of Ti alloys can be enhanced
seen on dragonfly wings. Such hydrothermally etched Ti nanowire by PGMs, and the corrosion resistance and passivation performance
surfaces with a hierarchical structure were found to have a higher can also be improved by inert alloying elements, such as Mo, Zr, Ta
wetting angle. Since bacteria on a nano biomimetic surface can be and Nb.
physically removed, surface hydrophobic modification can be used There are MIC and MICI phenomena on Ti. Colonization of organ-
as a non-toxic and environment-friendly approach in dealing with isms can occur on Ti surfaces, and the surface homogeneity is a
biofouling. significant factor influencing MIC. MICI can be attributed to the
A hydrophobic Ti surface is not only repellent to droplets, but formation of biofilm on the surface, which can act as barrier to sep-
also resistant to contaminative particles [221]. Thereupon, on such arate the Ti substrate from the environment media. In this case,
a surface biofilm-induced infections and marine biofouling can be oxygen reduction can also be influenced by microbial respiration.
effectively confined in theory. The improved corrosion resistance The fouling on Ti is much more severe than that on other metals
of Ti with a (super)hydrophobic surface can be ascribed to its water because of its biocompatible and inert properties. Controlling the
repellency, which can delay the initiation of corrosion. For exam- initial microbial adhesion is a crucial measure in antifouling. All the
ple, the superhydrophobic Ti surfaces anodized at 30 V followed existing antifouling strategies have some merits and drawbacks.
by dip-coating in 0.5 mol/L myristic acid were almost free from
microbial adhesion and thus largely reduced biofilm formation.
Meanwhile, corrosion resistance of Ti was noticeably increased 10. The future
after dip-coating [219]. Zhang et al. [218] found that the impedance
of Ti coated with low surface energy PTES compounds was much Generally speaking, Ti and its alloys are much more corrosion-
higher than that of anodized Ti. However, the corrosion resistance resistant than other engineering metals in marine environments.
resulting from surface hydrophobicity could not last long. The cor- With rapid development of today’s marine industry, Ti alloys will
rosion resistance of the PTES modified samples decreased with find increasingly wider and more applications in marine engi-
increasing immersion time, because the invasive ions could pen- neering. However, marine conditions can vary remarkably from
etrate the self-assembled monolayer (SAM) gradually through the summer to winter, from offshore to open ocean, and from equa-
defects in the coating and reacted with the metal substrate, and cor- tor to Arctic water and from splash zone to deep sea. Currently,
rosion of the substrate further destroyed the coating in return. The many interesting research areas have not been well exploited. It
groove-shaped structures on hydrophobic Ti surface were obtained is unclear if commercial Ti alloys are still corrosion resistant in
by immersion in H2 O2 (30%) followed by heating below 500 ◦ C. Cor- some extreme marine conditions. For example, in deep-sea where
rosion potentials of the electrodes became more positive, because mining and oil/gas production may be carried out, the environ-
a thick Ti oxide film formed on Ti acted as an electron barrier [222]. ment may become too harsh to commercially available Ti alloys.
Surface hydrophobic modification can be easily obtained, but Therefore, the main concern in the Ti industry will be whether exist-
its long-term effect on antifouling and anticorrosion needs to be ing Ti alloys can maintain the required properties under extreme
further enhanced in practical marine environments. marine conditions. Understanding the corrosion performance of
newly developed alloys in such extreme environments will also
undoubtedly be an interesting topic in corrosion science.
8.4. Other novel technologies
Biofouling is one of the critical challenges for Ti alloys as marine
engineering materials. Species derived from adhered organisms
Antifouling methods that could impel the dissolution and
may react with Ti alloys substrates. The micro-environment inside
desorption of bio-adhesives from fouling organisms have been
the fouling layer on Ti may be incredibly different from the outside
proposed, but the chemical-physical properties of adhesives and
in terms of the pH value, dissolved oxygen content, salinity metal
secretion from cellular processes must be understood before
ion and its concentration. It will be of significance to gain an insight
this technology could be further developed [166]. An achiev-
into the influence of metabolites of the adhered organisms on cor-
able, environment-friendly and cost-effective technology that can
rosion of Ti alloys in seawater. Such bio-corrosion is an important
reduce the adhesion of a biofouling layer on Ti and its alloys in
topic in corrosion science. It is anticipated that the bio-corrosion
marine environment is in high demand.
of Ti will continuously attract more and more corrosion scientists
and engineers in future, and the related biofouling behavior and
9. Summary mechanism will become important research topics.
There is always a surface oxide film on Ti and its alloys,
Owing to the protective oxide film, Ti exhibits excellent cor- and the film critically determines the corrosion or passivity
rosion resistance in many service environments. Currently, there of Ti and its alloys. The above-mentioned biofilms (biofouling
are many investigations on various aspects of the passive film on layers) are normally formed on the top of this surface film.

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 13

Thus, this film can affect the activity of organisms on its sur- and low-cost Ti alloys for marine engineering applications should
face. On the other hand, the bioactivity of the organisms can be prioritized in Ti industry.
also influence the formation and dissolution of the passive Obviously, development of low-cost, corrosion resistant and
oxide film. At present, the interaction between the Ti oxide anti-biofouling Ti alloys is a long-term solution to current prob-
film and biofouling layers has not been systematically investi- lems with Ti alloys in the marine industry. To meet some urgent
gated, and the synergistic effect of electrochemical reaction and challenges in marine engineering, short-term prevention mea-
organism’s activity on the degradation of Ti and its alloys in sures, such as cathodic protection, surface treatment/coating, and
seawater has not been comprehensively understood. A mechanis- antifouling techniques, are needed. Apart from the methods of mit-
tic study on these behavior may lead to an innovative solution igating passive film breakdown and organism growth on Ti alloy
to the severe biofouling problem with Ti alloys. Such funda- surfaces, a combination of individual anti-corrosion techniques and
mental research should be listed as a prioritized project for Ti anti-fouling methods may have an unexpected synergetic effect.
alloys. To the best of the authors’ knowledge, there is no publication on
The excellent corrosion resistance of Ti and its alloys is mainly the corrosion-fouling interaction for anodized Ti alloys. Hence, the
attributed to the presence of a passive film on their surfaces. effect of Ti alloy surface treatment on biofouling behavior could be
The measurement of the extremely thin (only several nanome- an interesting area of significance, which can be focused in future.
ters thick) film mainly consisting of TiO2 natively formed on the
surface in air is limited by the accuracy of existing techniques
Acknowledgements
today. Although some theoretical models have been proposed for
this passive film, its formation and damage mechanisms in marine
The beneficial discussion with Profs. Zhenhua Dan, Xiaogang Li,
environments have not been comprehensively verified. There is
Guangzhang Chen, Zhaoli Ma and Peng Ge during preparing this
an obvious gap between Ti passivity theory and experimental evi-
review paper is highly appreciated.
dence. Application of innovative or emerging surface techniques to
characterize the passive film nature in nano scale in situ will be key
to the comprehensive understanding the passive film. This will be References
a trend in Ti passivity investigations.
[1] C. Cui, B. Hu, L. Zhao, S. Liu, Mater. Des. 32 (2011) 1684–1691.
More specifically, the composition and microstructure of a [2] J.C. Williams, E.A. Starke, Acta Mater. 51 (2003) 5775–5799.
passive film are dependent on the substrate metal and the envi- [3] H.C. Flemming, Appl. Microbiol. Biotechnol. 59 (2002) 629–640.
ronmental factors. For a given Ti alloy, the species, such as Cl− , Br− , [4] F.H. Froes, H. Friedrich, J. Kiese, D. Bergoint, JOM 56 (2004) 40–44.
[5] R. Tejero, E. Anitua, G. Orive, Prog. Polym. Sci. 39 (2014) 1406–1447.
F− , I− , HCO3 − , Ca2+ , and dissolved O2 , from seawater may partici-
[6] F. Yu, O. Addison, S.J. Baker, A.J. Davenport, Int. J. Oral Sci. 7 (2015) 179–186.
pate in the formation or dissolution reactions of the passive film, [7] I.V. Gorynin, Mater. Sci. Eng A 263 (1999) 112–116.
penetrate into this film, or deposit on its surface. Therefore they can [8] J. Been, K. Faller, JOM 51 (1999) 21–24.
affect the passivation performance of such Ti alloys. The influence [9] H. Wake, H. Takahashi, T. Takimoto, H. Takayanagi, K. Ozawa, H. Kadoi, M.
Okochi, T. Matsunaga, Biotechnol. Bioeng. 95 (2006) 468–473.
of the species on passivity of Ti and its alloys has not been system- [10] E. Aragon, J. Woillez, C. Perice, F. Tabaries, M. Sitz, Mater. Des. 30 (2009)
atically investigated. A fundamental study on the effect of typical 1548–1555.
marine parameters on the passivity of Ti and its alloys must be [11] G. Gusmano, G. Montesperelli, G. Forte, E. Olzi, A. Benedetti, Desalination
183 (2005) 187–194.
carried out before Ti alloys are widely used as marine engineering [12] H. Richaudminier, H. Marchebois, P. Gerard, Mater. Technol. 24 (2009)
materials. 191–200.
The difficulty in comprehending Ti passivity or stability in [13] F. Mansfeld, G. Liu, H. Xiao, C.H. Tsai, B.J. Little, Corros. Sci. 36 (1994)
2063–2095.
marine environments lies in the complicated variation and unde- [14] J. Gopal, P. Muraleedharan, H. Sarvamangala, R.P. George, R.K. Dayal, B.V.R.
tectable interaction of the environmental factors. For example, the Tata, H.S. Khatak, K.A. Natarajan, Biofouling 24 (2008) 275–282.
environment temperature may be below the generally accepted [15] J.Q. Flores, Y.S. Joung, N.M. Kinsinger, X.L. Lu, C.R. Buie, S.L. Walker, Colloids
Surf. B 134 (2015) 204–212.
threshold for pitting or crevice corrosion of Ti alloys, but the sur- [16] B. Little, P. Wagner, F. Mansfeld, Int. Mater. Rev. 36 (1991) 253–272.
face oxide film can still break down, initiating crevice corrosion [17] C. Priya, G. Aravind, W.R. Thilagaraj, Bull. Mater. Sci. 39 (2016) 345–351.
and local acidification on the alloys, if there is bacterial activity or [18] R. Parrott, Eng. Fail. Anal. 44 (2014) 424–440.
[19] Z.G. Yang, Y. Gong, J.Z. Yuan, Mater. Corros. 63 (2012) 7–17.
metabolites involved in the process. Also, the oxide film may break
[20] L. Vedaprakash, R. Dineshram, K. Ratnam, K. Lakshmi, K. Jayaraj, S.M. Babu,
down due to mechanical damage due to SCC, corrosion fatigue, ero- R. Venkatesan, A. Shanmugam, Colloids Surf. B 106 (2013) 1–10.
sion, cavitation, or a stray current. Under some extreme conditions, [21] H.M. Shalaby, H. Al-Mazeedi, H. Gopal, N. Tanoli, Eng. Fail. Anal. 18 (2011)
e.g., in deep seawater with high pressure, low oxygen content and 1990–1997.
[22] G. Sternhell, P.D. Taylor, D. Itzhak, Corros. Rev. 20 (2002) 453–468.
low temperature, the thermodynamic and kinetic parameters of the [23] F. Hua, K. Mon, P. Pasupathi, G. Gordon, D. Shoesmith, JOM 57 (2005) 20–26.
corrosion on Ti and its alloys may deviate significantly from those [24] J. Pang, D.J. Blackwood, Corros. Sci. Technol. 14 (2015) 195–199.
in ordinary marine environments. Some estimated corrosion per- [25] M. Ishii, M. Kaneko, T. Oda, Nippon Steel Tech. Rep. 87 (2003) 49–56.
[26] B. Grosgogeat, L. Reclaru, M. Lissac, F. Dalard, Biomaterials 20 (1999)
formance based on common knowledge may become unreliable. It 933–941.
is quite possible that the corrosion of Ti and its alloys is more severe [27] K.L. Kilpadi, P.L. Chang, S.L. Bellis, J. Biomed. Mater. Res. 57 (2001) 258–267.
in real marine environments than in the laboratory. A survey of the [28] P. Giannoni, A. Muraglia, C. Giordano, R. Narcisi, R. Cancedda, R. Quarto, R.
Chiesa, Int. J. Artif. Organs 32 (2009) 811–820.
corrosion damage and relevant data acquisition for Ti and its alloys [29] H. Zitter, Mater. Corros. 39 (1988) 574–582.
in real marine environments will be meaningful and useful. [30] M. Simon, C. Lagneau, J. Moreno, M. Lissac, F. Dalard, B. Grosgogeat, Eur. J.
Apart from environmental factors, the composition and Oral Sci. 113 (2005) 537–545.
[31] Z.U. Rahman, L. Pompa, W. Haider, J. Mater. Eng. Perform. 24 (2015) 543.
microstructure of the substrate can significantly influence the [32] C.M. Bhadra, V.K. Truong, V.T.H. Pham, M. Al Kobaisi, G. Seniutinas, J.Y.
protectiveness of the passive film. Thereby, the effect of the sub- Wang, S. Juodkazis, R.J. Crawford, E.P. Ivanova, Sci. Rep. 5 (2015) 12.
strate as well as the processing conditions on the composition [33] K. Kondoh, N. Nakanishi, R. Taker, J. Umeda, General paper selections, TMS
2011 Supplemental Proceedings Vol 3 (2011) 93–100.
and micro-restructure of the passive film, and the passivation and
[34] J. Schmets, J.V. Muylder, M. Pourbaix, in: M. Pourbaix (Ed.), Titanium,
breakdown behaviors will always be an important interesting topic Pergamon Press, Oxford, 1966, pp. 213–222.
in Ti alloy studies, which may lead to development of a new alloy [35] A.M. Al-Mayouf, A.A. Al-Swayih, N.A. Al-Mobarak, A.S. Al-Jabab, Mater.
for marine engineering applications. Since there is a concern about Chem. Phys. 86 (2004) 320–329.
[36] D.-S. Kong, Langmuir 24 (2008) 5324–5331.
the survival of existing commercial alloys in extreme marine envi- [37] N. Casillas, S.R. Snyder, W.H. Smyrl, H.S. White, J. Phys. Chem. 95 (1991)
ronments, development of new corrosion-resistant, anti-biofouling 7002–7007.

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
14 S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx

[38] R.A. Piggott, L.L. Shreir, Nature 189 (1961) 216–217. [92] H.H. Huang, Biomaterials 24 (2003) 275–282.
[39] H.B. Bomberger, P.J. Cambourelis, G.E. Hutchinson, J. Electrochem. Soc. 101 [93] M. Nakagawa, S. Matsuya, T. Shiraishi, M. Ohta, J. Dent. Res. 78 (1999)
(1954) 442–447. 1568–1572.
[40] E. Eisenbarth, D. Velten, M. Muller, R. Thull, J. Breme, Biomaterials 25 (2004) [94] Z.B. Wang, H.X. Hu, Y.G. Zheng, Electrochim. Acta 170 (2015) 300–310.
5705–5713. [95] C.S. Brossia, G.A. Cragnolino, Corros. Sci. 46 (2004) 1693–1711.
[41] C. Vasilescu, S.I. Drob, P. Osiceanu, J.M. Calderon-Moreno, P. Drob, E. [96] T.R. Beck, J. Electrochem. Soc. 120 (1973) 1310–1316.
Vasilescu, Mater. Corros. 66 (2015) 971–981. [97] R. Ittah, E. Amsellem, D. Itzhak, Int. J. Electrochem. Sci. 9 (2014) 633–643.
[42] M. Geetha, U.K. Mudali, A.K. Gogia, R. Asokamani, B. Raj, Corros. Sci. 46 [98] D. Itzhak, T. Greenberg, Mater. Sci. Eng. A 302 (2001) 135–140.
(2004) 877–892. [99] S.Z. Huo, X.X. Meng, Corros. Sci. 31 (1990) 281–286.
[43] E. McCafferty, J.P. Wightman, Appl. Surf. Sci. 143 (1999) 92–100. [100] D. Sazou, K. Saltidou, M. Pagitsas, Electrochim. Acta 76 (2012) 48–61.
[44] C.J. Boxley, H.S. White, C.E. Gardner, J.V. Macpherson, J. Phys. Chem. B 107 [101] A.M. Fekry, Electrochim. Acta 54 (2009) 3480–3489.
(2003) 9677–9680. [102] J. Vaughan, A. Alfantazi, J. Electrochem. Soc. 153 (2006) B6–B12.
[45] A.F. Carley, P.R. Chalker, J.C. Riviere, M.W. Roberts, J. Chem. Soc. Faraday [103] I. Dugdale, J.B. Cotton, Corros. Sci. 4 (1964) 397–411.
Trans. 83 (1987) 351–370. [104] N.T. Thomas, K. Nobe, J. Electrochem. Soc. 119 (1969) 1450.
[46] T. Ohtsuka, M. Masuda, N. Sato, J. Electrochem. Soc. 132 (1985) 787–792. [105] J. Liu, A. Alfantazi, E. Asselin, J. Electrochem. Soc. 162 (2015) C189–C196.
[47] J. Lausmaa, B. Kasemo, H. Mattsson, Appl. Surf. Sci. 44 (1990) 133–146. [106] J. Liu, A. Alfantazi, E. Asselin, J. Electrochem. Soc. 162 (2015) E289–E295.
[48] H.Y.H. Chan, S.Z. Zou, M.J. Weaver, J. Phys. Chem. B 103 (1999) 11141–11151. [107] I.N. Andijani, S. Ahmad, A.U. Malik, Desalination 129 (2000) 45–51.
[49] C.Y. Chao, L.F. Lin, D.D. Macdonald, J. Electrochem. Soc. 128 (1981) [108] H. Satoh, Bull. Soc Sea Water Sci. 44 (1990) 200–208.
1187–1194. [109] M.G. Fontana, N.D. Greene, Corrosion Engineering (McGraw-Hill Series in
[50] D.D. Macdonald, S.R. Biaggio, H.K. Song, J. Electrochem. Soc. 139 (1992) Materials Science and Engineering), Macgraw-Hill International, New York,
170–177. 1978.
[51] G. Song, Corros. Sci. 47 (2005) 1953–1987. [110] N. Rajendran, T. Nishimura, Mater. Corros. 58 (2007) 334–339.
[52] L. Zhang, D.D. Macdonald, E. Sikora, J. Sikora, J. Electrochem. Soc. 145 (1998) [111] A.J. Hatch, H.W. Rosenberg, E.F. Erbin, Am. Soc. Testing Mats. (1966) 122.
898–905. [112] C. Mulyana, R. Adiprana, A.H. Saad, H.M. Ridwan, F. Muhammad, in: I.M. Joni,
[53] D.D. Macdonald, Electrochim. Acta 56 (2011) 1761–1772. C. Panatarani (Eds.), 2nd Padjadjaran International Physics Symposium
[54] C. Wagner, Ber. Bunsen Ges. 63 (2015) 772–782. 2015, Amer Inst Physics Melville, 2016.
[55] D.D. Macdonald, Russ. J. Electrochem. 48 (2012) 235–258. [113] W.L. Wheatfall, Nav. Eng. J. 79 (1967) 611–618.
[56] A. Seyeux, V. Maurice, P. Marcus, J. Electrochem. Soc. 160 (2013) C189–C196. [114] M.A. Pelensky, J.J. Jaworski, A. Gallaccio, Air, Soil and Sea Galvanic Corrosion
[57] B. Krishnamurthy, R.E. White, H.J. Ploehn, Electrochim. Acta 47 (2002) Investigation at Panama Canal Zone, ASTM STP576, 1976, pp. 94.
2505–2513. [115] F.M. Reinhart, Titanium and Titanium Alloys, Tech. Note N-921, U.S. Naval
[58] N. Cabrera, N.F. Mott, Rep. Prog. Phys. 7 (1949) 163–184. Civil Engineering Lab, Port Hueneme, California, 1967 (Sept).
[59] M.M. Lohrengel, Mater. Sci. Eng R 11 (1993) 243–294. [116] H. Sun, L. Liu, Y. Li, L. Ma, Y. Yan, Corros. Sci. 77 (2013) 77–87.
[60] S. Tanaka, Y. Fukushima, I. Nakamura, T. Tanaki, G. Jerkiewicz, ACS Appl. [117] J. Pang, D.J. Blackwood, Corros. Sci. 105 (2016) 17–24.
Mat. Interfaces 5 (2013) 3340–3347. [118] X. Xin, J. Xue, M. Dong, Protection Corrosion, Application of Titanium, Anhui
[61] J.L. Delplancke, A. Garnier, Y. Massiani, R. Winand, Electrochim. Acta 39 Science and Technology Press, Hefei, 1988.
(1994) 1281–1289. [119] J.M.A. Kader, F.M.A. Wahab, H.A. Shayeb, M.G.A. Khedr, Br. Corros. J. 16
[62] X. He, J.J. NoëL, D.W. Shoesmith, J. Electrochem. Soc. 149 (2002) B440–B449. (1981) 111–114.
[63] L. Sun, S. Zhang, X.W. Sun, X. He, J. Electroanal. Chem. 637 (2009) 6–12. [120] A.M.S. Eldin, A.A. Hammoud, Thin Solid Films 167 (1988) 269–280.
[64] K. Fushimi, T. Okawa, K. Azumi, M. Seo, J. Electrochem. Soc. 147 (2000) [121] A.M.S. ElDin, T.M.H. Saber, A.M.T. ElDin, Desalination 107 (1996) 265–276.
524–529. [122] X.Q. Du, Q.S. Yang, Y. Chen, Y. Yang, Z. Zhang, Trans. Nonferrous Met. Soc.
[65] A. Müller, A. Benninghoven, Surf. Sci. 41 (1974) 493–503. China 24 (2014) 570–581.
[66] L. Johansson, A. Hagstrom, A. Platau, S. Karlsson, Phys. Status Solidi B 83 [123] T. Greenberg, D. Itzhak, Corros. Rev. 23 (2005) 405–413.
(1977) 77–84. [124] M.E. El-Dahshan, A.M.S. El Din, H.H. Haggag, Desalination 142 (2002)
[67] J. Pouilleau, D. Devilliers, F. Garrido, S. DurandVidal, Mater. Sci. Eng B 47 161–169.
(1997) 235–243. [125] C.L. Wang, Q.F. Li, J.H. Wu, in: Q.F. Li, Y.L. Li, M.H. Aliabadi (Eds.), Advances In
[68] K. Azumi, M. Seo, Corros. Sci. 43 (2001) 533–546. Fracture And Damage Mechanics XI, 2013, pp. 325–328.
[69] S.V. Gnedenkov, P.S. Gordienko, S.L. Sinebrukhov, O.A. Khrisanphova, T.M. [126] S.A. Namjoshi, S. Mall, Int. J. Fatigue 23 (2001) 455–461.
Skorobogatova, Corrosion 56 (2000) 24–31. [127] A. Hutson, C. Neslen, T. Nicholas, Tribol. Int. 36 (2003) 133–143.
[70] T. Ohtsuka, T. Otsuki, Corros. Sci. 40 (1998) 951–958. [128] L.C. Lietch, H. Lee, S. Mall, Mater. Sci. Eng A 403 (2005) 281–289.
[71] J. Li, S.J. Li, Y.L. Hao, R. Yang, Int. J. Hydrogen Energy 39 (2014) 17452–17459. [129] H.Y. Ding, Z.D. Dai, Rare Met. Mater. Eng. 36 (2007) 778–781.
[72] S. Wang, L. Pan, J.-J. Song, W. Mi, J.-J. Zou, L. Wang, X. Zhang, J. Am. Chem. [130] C.L. Briant, Z.F. Wang, N. Chollocoop, Corros. Sci. 44 (2002) 1875–1888.
Soc. 137 (2015) 2975–2983. [131] S.K. Yen, Corros. Sci. 41 (1999) 2031–2051.
[73] M.K. Nowotny, P. Bogdanoff, T. Dittrich, S. Fiechter, A. Fujishima, H. [132] K. Azumi, M. Seo, Corros. Sci. 47 (2005) 2470–2476.
Tributsch, Mater. Lett. 64 (2010) 928–930. [133] D.K. Peacock, F. Corr, Corros. Rev. 18 (2000) 295–330.
[74] P. Ducheyne, S. Radin, M. Heughebaert, J.C. Heughebaert, Biomaterials 11 [134] W.T. Tsai, C.P. Ju, Y.N. Wen, J.T. Lee, Surf. Coat. Technol. 31 (1987) 401–408.
(1990) 244–254. [135] H.M. Shalaby, Eng. Fail. Anal. 13 (2006) 780–788.
[75] M. Gottlander, C.B. Johansson, A. Wennerberg, T. Albrektsson, S. Radin, P. [136] S.B. Farina, G.S. Duffo, Corrosion 63 (2007) 450–461.
Ducheyne, Biomaterials 18 (1997) 551–557. [137] D.M. Kocak, H. Painter, G. Anderson, Ieee Oceans (2011) 7.
[76] V.N. Bagratashvili, E.N. Antonov, E.N. Sobol, V.K. Popov, S.M. Howdle, Appl. [138] H. Mochizuki, M. Yokota, S. Hattori, Wear 262 (2007) 522–528.
Phys. Lett. 66 (1995) 2451–2453. [139] R. Murakami, W.G. Ferguson, Fatigue Fract. Eng. Mater. Struct. 16 (1993)
[77] P. Ducheyne, W.V. Raemdonck, J.C. Heughebaert, M. Heughebaert, 255–265.
Biomaterials 7 (1986) 97–103. [140] V.A. Shulov, N.A. Nochovnaia, G.E. Remnev, Surf. Coat. Technol. 158-159
[78] J. Lee, H. Aoki, Bio-Med. Mater. Eng. 5 (1995) 49–58. (2002) 488–493.
[79] M. Catauro, S.P. Nunziante, F. Papale, F. Bollino, J. Sol-Gel Sci. Technol. 76 [141] A.J. Sedriks, Corrosion 31 (2013) 60–65.
(2015) 241–250. [142] G.Y. Zhang, L.N. Yang, H. Zhang, J.J. Wu, Acta Phys. Sin. 59 (2010) 2022–2026
[80] Z.J. Liu, X.H. Liu, U. Donatus, G.E. Thompson, P. Skeldon, Int. J. Electrochem. (in Chinese).
Sci. 9 (2014) 3558–3573. [143] F. Cardarelli, Materials Handbook? A Concise Desktop Reference, Springer,
[81] M.E.P. Souza, M. Ballester, C.M.A. Freire, Surf. Coat. Technol. 201 (2007) London, 2000, pp. 302–319, ISBN 978–144-7136–7484.
7775–7780. [144] C.S. Brossia, G.A. Cragnolino, Corrosion 57 (2001) 768–776.
[82] X. Huang, Z. Liu, Surf. Coat. Technol. 232 (2013) 224–233. [145] R.F. Sandenbergh, E. Van der Lingen, Corros. Sci. 47 (2005) 3300–3311.
[83] N. Masahashi, S. Semboshi, N. Ohtsu, M. Oku, Thin Solid Films 516 (2008) [146] F. Hua, K. Mon, P. Pasupathi, G. Gordon, D. Shoesmith, Corrosion 61 (2005)
7488–7496. 987–1003.
[84] S.B. Basame, H.S. White, J. Electrochem. Soc. 147 (2000) 1376–1381. [147] M.A. Langoy, S.R. Stock, Metall. Mater. Trans. A 32 (2001) 2297–2314.
[85] N. Casillas, S.J. Charlebois, W.H. Smyrl, H.S. White, J. Electrochem. Soc. 141 [148] R.W. Schutz, Platinum Met. Rev. 40 (1996) 54–61.
(1994) 59–60. [149] R.W. Schutz, J. Mol. Catal. A: Chem. 396 (1992) 328–334.
[86] K.S.E. Al-Malahy, T. Hodgkiess, Desalination 158 (2003) 35–42. [150] R.W. Schutz, Corrosion 59 (2003) 1043–1057.
[87] D. Liu, Corrosion and Protection of Material, Northwestern Polytechnical [151] D.G. Kolman, J.R. Scully, J. Electrochem. Soc. 140 (1993) 2771–2779.
University Press, Xi’an, 2006 (in Chinese). [152] H. Habazaki, M. Uozumi, H. Konno, K. Shimizu, S. Nagata, K. Asami, P.
[88] R.W. Schutz, J.S. Grauman, in: C.S. Young, J.C. Durham (Eds.), Industrial Skeldon, G.E. Thompson, Electrochim. Acta 47 (2002) 3837–3845.
Applications of Titanium and Zirconium, Vol. 4, 1986, pp. 130–143 (ASTM [153] J.E.G. González, J.C. Mirza-Rosca, J. Biomed. Mater. Res. 471 (1999) 109–115.
STP917). [154] R.C. Newman, Corros. Sci. 16 (1985) 341–350.
[89] The Japan Titanium Society, in: L.Z. Zhou (Ed.), Trans. Titanium Materials [155] F. Qin, X. Wang, A. Kawashima, S. Zhu, H. Kimura, A. Inoue, Mater. Trans. 47
and Applications, Metallurgical Industry Press, Beijing, 2008. (2006) 1934–1937.
[90] H. Satoh, F. Kamikubo, K. Shimogori, T. Fukuzuka, Zairyo-to-Kankyo 32 [156] B.L. Wang, Y.F. Zheng, L.C. Zhao, Mater. Corros. 60 (2009) 788–794.
(1983) 69–75. [157] S. Baltatu, P. Vizureanu, D. Mareci, L.C. Burtan, C. Chiruta, L.C. Trinca, Mater.
[91] H. Park, W. Choi, J. Phys. Chem. B 108 (2004) 4086–4093. Corros. 67 (2016) 1314–1320.

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021
G Model
JMST-1111; No. of Pages 15 ARTICLE IN PRESS
S. Yan et al. / Journal of Materials Science & Technology xxx (2017) xxx–xxx 15

[158] Y. Tanaka, M. Nakai, T. Akahori, M. Niinomi, Y. Tsutsumi, H. Doi, T. Hanawa, [192] N. Washizu, Y. Katada, T. Kodama, Corros. Sci. 46 (2004) 1291–1300.
Corros. Sci. 50 (2008) 2111–2116. [193] K. Xu, S.C. Dexter, G.W. Luther, Corrosion 54 (1998) 814–823.
[159] R.J. Hanrahan, D.P. Butt, Oxid. Met. 47 (1997) 317–353. [194] C. Sun, J. Xu, F. Wang, Ind. Eng. Chem. Res. 50 (2011) 12797–12806.
[160] Y.S. Chen, C.J. Rosa, Oxid. Met. 14 (1980) 167–185. [195] M.J. Dempsey, Bot. Mar. 24 (1981) 185–191.
[161] M. Metikos-Hukovic, A. Kwokal, J. Piljac, Biomaterials 24 (2003) 3765–3775. [196] A.G. Nurioglu, A.C.C. Esteves, G. de With, J. Mater. Chem. B 3 (2015)
[162] Y.Z. Liu, X.T. Zu, C. Li, S.Y. Qiu, X.Q. Huang, L.M. Wang, Corros. Sci. 49 (2007) 6547–6570.
1069–1080. [197] H. Anwar, M.K. Dasgupta, J.W. Costerton, Antimicrob. Agents Chemother. 34
[163] M.F. Semlitsch, H. Weber, R.M. Streicher, R. Schon, Biomaterials 13 (1992) (1990) 2043–2046.
781–788. [198] D. Debeer, R. Srinivasan, P.S. Stewart, Appl. Environ. Microbiol. 60 (1994)
[164] I. Milosev, T. Kosec, H.H. Strehblow, Electrochim. Acta 53 (2008) 3547–3558. 4339–4344.
[165] P.K.A. Azis, I. Al-Tisan, N. Sasikumar, Desalination 135 (2001) 69–82. [199] J.R. Henschel, P.A. Cook, Biofouling 2 (1990) 1–11.
[166] M.E. Callow, J.E. Callow, Biologist 49 (2002) 10–14. [200] V. Vishwakarma, J. Josephine, R.P. George, R. Krishnan, S. Dash, M.
[167] T.B. Buzovkina, V.A. Aleksandrov, L.I. Shlyaga, N.D. Perekhvalskaya, Prot. Kamruddin, S. Kalavathi, N. Manoharan, A.K. Tyagi, R.K. Dayal, Biofouling 25
Met. 21 (1985) 657–659. (2009) 705–710.
[168] R.O. Lewis, Mater. Perform. 21 (1982) 31–38. [201] M.B. Mcneil, J.M. Jones, B.J. Little, Corrosion 47 (1991) 674–677.
[169] R.B. Griffin, L.R. Cornwell, W. Seitz, E. Estes, Mater. Perform. 28 (1989) 71–74. [202] G. Borkow, J. Gabbay, Curr. Med. Chem. 12 (2005) 2163–2175.
[170] X.X. Sheng, Y.P. Ting, S.O. Pehkonen, Corros. Sci. 49 (2007) 2159–2176. [203] A.P. Negri, A.J. Heyward, Mar. Environ. Res. 51 (2001) 17–27.
[171] A. Nagiub, F. Mansfeld, Electrochim. Acta 47 (2002) 2319–2333. [204] R.L. Townsin, Biofouling 19 (Suppl) (2003) 9–15.
[172] R. Javaherdashti, Appl. Microbiol. Biotechnol. 91 (2011) 1507–1517. [205] H.X. Hong, Y.S. Yin, H.F. Wang, S.G. Chen, T. Liu, L.L. Zhao, in: Y.S. Yin, X.
[173] F. Liu, J. Zhang, C. Sun, Z. Yu, B. Hou, Corros. Sci. 83 (2014) 375–381. Wang (Eds.), Multi-Functional Materials And Structures, Part 1–2, 2009, pp.
[174] R. Javaherdashti, Anti-Corros. Methods Mater. 46 (1999) 173–180. 1103–1106.
[175] D.W. Zhang, F.C. Zhou, K. Xiao, T.Y. Cui, H.C. Qian, X.G. Li, J. Mater. Eng. [206] T. Matsunaga, Y. Namba, T. Nakajima, Bioelectrochem. Bioenerg. 13 (1984)
Perform. 24 (2015) 2688–2698. 393–400.
[176] T.S. Rao, A.J. Kora, B. Anupkumar, S.V. Narasimhan, R. Feser, Corros. Sci. 47 [207] M. Okochi, T.K. Lim, N. Nakamura, T. Matsunaga, Appl. Microbiol. Biotechnol.
(2005) 1071–1084. 47 (1997) 18–22.
[177] B. Anandkumar, N.G. Krishna, R.P. George, N. Parvathavarthini, U.K. Mudali, [208] M. Okochi, N. Nakamura, T. Matsunaga, Clean Prod. Processes 1 (1998)
J. Control Sci. Eng. 16 (2013) (ISSN 1466–8858). 53–59.
[178] R. Javaherdashti, Microbiologically Influenced Corrosion: An Engineering [209] T. Matsunaga, Y. Namba, Anal. Chem. 56 (1984) 798–801.
Insight, in: B. Derby (Ed.), Springer, London, 2008, pp. 29–71. [210] M.T. Ehrensberger, M.E. Tobias, S.R. Nodzo, L.A. Hansen, N.R. Luke-Marshall,
[179] M. Dubiel, C.H. Hsu, C.C. Chien, F. Mansfeld, D.K. Newman, Appl. Environ. R.F. Cole, L.M. Wild, A.A. Campagnari, Biomaterials 41 (2015) 97–105.
Microbiol. 68 (2002) 1440–1445. [211] J. Gopal, R.P. George, P. Muraleedharan, S. Kalavathi, S. Banerjee, R.K. Dayal,
[180] M.A. Ameer, A.M. Fekry, S.M. Shanab, Int. J. Electrochem. Sci. 6 (2011) H.S. Khatak, J. Mater. Sci. 42 (2007) 5152–5158.
1572–1585. [212] R. Blossey, Nat. Mater. 2 (2003) 301–306.
[181] R.A. Brizzolara, D.J. Nordham, M. Walch, R.M. Lennen, R. Simmons, E. [213] B. Liu, W.F. Xu, H. Li, L. Liu, Y.H. Shen, A.J. Xie, Chem J. Chinese U 34 (2013)
Burnett, M.S. Mazzola, Biofouling 19 (2003) 19–35. 2191–2195.
[182] H.M. Shalaby, H. Al-Mazeedi, H. Gopal, N. Tanoli, Eng. Fail. Anal. 18 (2011) [214] T. Sun, L. Feng, X. Gao, L. Jiang, Acc. Chem. Res. 38 (2005) 644–652.
1990–1997. [215] G. Rølla, J.E. Ellingsen, B. Herlofson, Biofouling 3 (1991) 175–181.
[183] K. Nagata, K. Sudo, A. Kawabe, I. Katsuyama, Sumitomo Light Met. Tech. Rep. [216] M. Quirynen, H.C.V.D. Mei, C.M.L. Bollen, G.I. Geertsema-Doornbusch, H.J.
27 (1986) 80–90. Busscher, D.V. Steenberghe, Colloids Surf. B 2 (1994) 25–31.
[184] S. Palraj, G. Venkatachari, Desalination 230 (2008) 92–99. [217] E. Everaert, H.F. Mahieu, B. van de Belt-Gritter, A. Peeters, G.J. Verkerke, H.C.
[185] N. Bellou, F. Colijn, E. Papathanassiou, Nucl. Instrum. Methods Phys. Res van der Mei, H.J. Busscher, Arch. Otolaryngol Head Neck Surg. 125 (1999)
Sect. A 626 (2011) S102–S105. 1329–1332.
[186] R.E. Perez-Roa, M.A. Anderson, D. Rittschof, B. Orihuela, D. Wendt, G.L. [218] F. Zhang, S.G. Chen, L.H. Dong, Y.H. Lei, T. Liu, Y.S. Yin, Appl. Surf. Sci. 257
Kowalke, D.R. Noguera, Biofouling 24 (2008) 177–184. (2011) 2587–2591.
[187] V. L’Hostis, C. Dagbert, D. Feron, Electrochim. Acta 48 (2003) 1451–1458. [219] P.V. Mahalakshmi, S.C. Vanithakumari, J. Gopal, U.K. Mudali, B. Raj, Curr. Sci.
[188] J. Landoulsi, K.E. Cooksey, V. Dupres, Biofouling 27 (2011) 1105–1124. 101 (2011) 1328–1336.
[189] B. Anandkumar, R.P. George, K. Kamaraj, N. Parvathavarthini, U.K. Mudali, T. [220] D.V. Bavykin, F.C. Walsh, Eur. J. Inorg. Chem. 40 (2009) 977–997.
Indian I. Metals 70 (2017) 1075–1081. [221] C. Neinhuis, W. Barthlott, Ann. Bot. 79 (1997) 667–677.
[190] J.P. Busalmen, M. Vazquez, S.R. de Sanchez, Electrochim. Acta 47 (2002) [222] M. Wang, W. Wang, B.L. He, M.L. Sun, Y.S. Bao, Y.S. Yin, L. Liu, W.Y. Zou, X.F.
1857–1865. Xu, Mater. Corros. 62 (2011) 320–325.
[191] W.H. Dickinson, Z. Lewandowski, R.D. Geer, Corrosion 52 (1996) 910–920.

Please cite this article in press as: S. Yan, et al., J. Mater. Sci. Technol. (2017), https://doi.org/10.1016/j.jmst.2017.11.021

You might also like