Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Chemistry and Physics 233 (2019) 390–398

Contents lists available at ScienceDirect

Materials Chemistry and Physics


journal homepage: www.elsevier.com/locate/matchemphys

Study of electrochemical corrosion of biocompatible Co–Cr and Ni–Cr


dental alloys in artificial saliva. Influence of pH of the solution
Graziella L. Turdean a, *, Antarinia Craciun b, Daniela Popa b, Mariana Constantiniuc b, **
a
“Babes-Bolyai” University, Faculty of Chemistry and Chemical Engineering, Chemical Engineering Department, Research Center of Electrochemistry and
Nonconventional Materials, 11 Arany Janos St., Cluj-Napoca, Romania
b
“Iuliu Hatieganu” University of Medicine and Pharmacy, Faculty of Dental Medicine, 8 Victor Babes St., Cluj-Napoca, Romania

H I G H L I G H T S

� The resistance to corrosion of alloys is due to formation of a passive oxide film.


� The OCP, LP, Tafel plots and EIS spectra are in accordance with SEM and EDXS.
� From EIS results Co–Cr is more resistant to corrosion in artificial saliva of pH 3.

A R T I C L E I N F O A B S T R A C T

Keywords: The pH of saliva is changeable and variable in the oral environment in function of the quality of foods and liquids
Biomaterials ingested. Also, the biocompatibility of a material for dental applications is checked not only by clinical and
Corrosion epidemiological studies, but also by corrosion resistance studies. The aim of this study was to investigate the
Electrochemical impedance spectroscopy
influence of the pH value of artificial saliva on the corrosion of the Co–Cr and Ni–Cr dental alloys, using different
electrochemical techniques such as: cyclic voltammetry, open circuit potential (OCP) measurement, polarization
curves and electrochemical impedance spectroscopy (EIS). The obtained results are complementary with scan­
ning electron microscopy (SEM) coupled with energy-dispersive X-ray spectroscopy (EDXS) analysis, proving
that Co–Cr alloy is more resistant in time, and recommending it for successful treatment of patients with dental
prosthesis having metal frameworks.

1. Introduction with the Ni–Cr alloys from 1930s [2–5]. Metallic materials for applica­
tions in dentistry have special demands including: biocompatibility,
The most accepted definition of the biomaterial is: “any substance or non-toxicity, stability, safety and mechanical properties (as mechanical
combination of substances (other than drugs), synthetic or natural in strength, wear resistance, ductility for obtaining desired thinner shapes
origin, which can be used for any period of time, which enhances or in oral cavity) [4,6,7]. Among these properties, the biocompatibility is
replaces partially or totally any tissue, organ or function of the body, in the most important requirement because the metals/alloys being in
order to maintain or improve the quality of life of an individual” [1]. close contact with oral tissues could lead to adverse tissue reactions
Thus, the biomaterials play a significant role in improving the life and/or hypersensitivity reactions [8,9].
quality and for dental medical applications they could be: metals/alloys, During time, various metals have been investigated as implants, such
ceramics, polymers and composites. In the field of prosthetic dentistry as: aluminum, copper, zinc, iron and carbon steels, silver, nickel, and
where the metallic biomaterials are widely used, from the beginning of magnesium [10]. Despite the restrictive legislation imposed by the EU
19th century, the expensive precious or noble metal alloys (Au, Pt, Pd for the protection of the population [3,11,12] and despite the existence
and Ag) were replaced by “non precious” alloys (Ni and Co based), first of a great diversity of accessible metals/alloys on the materials market,
with the cobalt-chromium alloy patented in 1907 by Haynes, and then, only a few of them can be used in the fabrication of metal frameworks of

* Corresponding author.
** Corresponding author.
E-mail addresses: gturdean@chem.ubbcluj.ro (G.L. Turdean), mconstantiniuc@umfcluj.ro (M. Constantiniuc).

https://doi.org/10.1016/j.matchemphys.2019.05.041
Received 19 February 2019; Received in revised form 14 May 2019; Accepted 16 May 2019
Available online 17 May 2019
0254-0584/© 2019 Published by Elsevier B.V.
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Fig. 1. Cyclic voltammograms at Co–Cr (A) and Ni–Cr (B) alloys interfaces in artificial saliva solution of pH 3 (red line) and pH 7 (black line). Experimental
conditions: scan rate, 0.05 V s 1; starting potential, 1 V vs. Ag/AgCl, KClsat; forward scan (continue line), reverse scan (dot line). (For interpretation of the references
to colour in this figure legend, the reader is referred to the Web version of this article.)

partial removable dental prosthesis and for fixed prosthesis [13]. Such 2. Experimental
materials which can respond to the nowadays requirements of dental
medicine are: 316L stainless steel (316L SS) [5], cobalt-based (Co–Cr) 2.1. Materials and reagents
alloys [4,7,14–18], nickel-based (Ni–Cr) alloys [3,15,17,19], titanium
and its alloys [5,8,20,21], and “smart materials” as shape memory alloys The Ni–Cr (Ni 63%, Cr 25%, Mo 10%, Si 1.5%, others Mn, Al, Zr, Ce,
(i.e., Ni–Ti alloy called Nitinol) [5,6]. La) and Co–Cr (Co 65%, Cr 29%, Mo 5%, C 0.4%, Si 0.35%, Mn 0.25%)
A good partial removable dental prosthesis having metal frameworks dental metallic alloys were purchased by MESA di Sala Giacomo & C S.n.
which avoid producing hypersensitivity responses or adverse tissue re­ c. The samples were cut and embedded into a Teflon cylinder such that a
actions must have an improved anticorrosive resistance due to a spon­ 0.07065 cm2 active surface could be investigated. Because natural saliva
taneous formation on their surfaces, in air or in electrolytes at open is a really complex matrix, containing a great variety of compounds, and
circuit, of a thin oxide film [22]. Usually, the role of the passive oxide its properties are very changeable and unstable (during the time of day
layers is to act as a barrier to electron flow between the metal/alloy and or due to food consumption), it is quite impossible to obtain an artificial
the saliva, although the aggressive environment due to saliva solution of saliva solution which replicates it. From no fewer than 60 formulae
pH 3 could react with the surface elements, thus changing or causing known today, it was chosen a receipt proposed by Mondelli [27], which
defects in the homogeneity passive oxide layers of the alloys. Conse­ was prepared by using appropriate quantities of the following salts: NaCl
quently, is necessary a critical and deep analysis of the corrosion resis­ (0.5 g/L), KCl (0.5 g/L), CaCl2*2H2O (0.795 g/L), NaH2PO4*H2O
tance of a dental metallic alloy, which depends on the composition, (0.78 g/L), urea (1 g/L), (NH4)2SO4 (0.3 g/L), NaHCO3 (0.1 g/L) sup­
structure and properties of the thin passive oxide film [23]. Under plied from Sigma Aldrich and “Reactivul” Bucuresti. Diluted H3PO4
physiologic conditions, the undamaged integrity of the passive oxide (Sigma Aldrich) was used for pH adjustment. Distilled water was used
film could be correlated with the mechanical and chemical stability of for preparing all solutions. All chemicals having analytical purity were
the alloy [8,9]. Although in the research reports there are evidences that used as received without any purification.
the passive oxide film increases in passivity over time, a question
waiting response is: “how many metallic ions are released and swal­
2.2. Electrochemical methods
lowed by a patient having partial metallic alloy prosthesis, during its life
[24]? And how these corrosive behavior (i.e., attrition–corrosion) is
Electrochemical measurements were performed with an AutoLab
influenced by the variation of saliva’s pH [6,20] due to the food intakes
potentiostat (PGSTAT302 N EcoChemie, Utrecht, Netherlands)
(acidic food and beverage, dietary supplements, drinking water or tea)
controlled by GPES 4.7 software for cyclic/linear voltammetry and open
or the used of toothpastes, mouth rinses, and/or orthodontic gels [4,8,9,
circuit potential measurements, and FRA 2.1 software for impedance
24,25]?”
measurements. A conventional three-electrode cell, equipped with a
In order to evaluate the long-term successful and safety treatment of
working electrode of Ni–Cr and Co–Cr alloys, a Ag/AgCl, KClsat refer­
patients [8–10,26], and bearing in mind the very large use of metallic
ence electrode and a platinum wire auxiliary electrode, was used for
alloys, despite the fact that new polymeric or composite materials exist
electrochemical characterization of the dental metallic alloy electrodes.
on the market, a detailed corrosion resistance study is important and
Cyclic voltammetry measurements were carried out on non-
very interesting from statistical point of view on the impact of using such
stabilized electrodes using a potential scan rate of 0.050 Vs-1.
kind of materials on the population health in a certain geographical
For corrosion tests, prior to measurements, the metallic alloys were
region dependent on an unique dental materials market. The aim of the
stabilized under open circuit conditions for 600 s, 3600 s or 9000 s, and
present study was to evaluate the influence of the pH of artificial saliva
the open circuit potential (OCP) was measured. The linear polarization
on the corrosion of two commercially metallic alloys, namely Ni–Cr and
resistance (LPR, Mansfeld curves [28]) measurements were performed
Co–Cr alloys, available on Romanian dental material market. For the
by polarizing the electrodes in a potential range of �20 mV vs. OCP,
first time to our knowledge, the corrosion resistance was evaluated in
allowing the estimation of polarization resistance (Rp) values from the
biomimetic oral cavity environment, by cyclic voltammetry, open circuit
slope of the fitted I – E plot. The polarization curves were recorded in a
potential (OCP), polarization curves and electrochemical impedance
potential range of �200 mV vs. OCP, in order to estimate icorr and Ecorr
spectroscopy (EIS).
by using the Tafel extrapolation method.
Impedance spectra at dental Ni–Cr and Co–Cr alloys interface were
recorded every hour, within 24 h, at the OCP value, by using an AC

391
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Fig. 2. Dependence of open circuit potential (OCP) over time for the Ni–Cr
alloy immersed in artificial saliva solution of pH 3.
Fig. 3. Linear polarization curves for Ni–Cr alloy at different time for stabili­
zation of OCP. Experimental conditions: electrolyte, artificial saliva solution
signal with amplitude of �10 mV and a frequency interval from 10 kHz
(pH 3), scan rate, 0.5 mV s 1; interval of potential, � 0.020 V vs. OCP.
down to 0.01 Hz. EIS data were plotted as Nyquist and Bode diagrams.
Using an equivalent electric circuit, the experimental impedance data
were fitted using ZSimpWin 3.21 software, the values of circuit com­ alloys interface recorded in the potential domain of 1 to 1 V vs. Ag/
ponents were subsequently correlated with the processes occurring in AgCl, KClsat, in artificial saliva solutions having two different pH values
the electrochemical cell. (i.e., 3 and 7). Both alloys show a negative hysteresis at potentials
The pH of the electrolyte solutions was adjusted using a glass com­ greater than 0.8 V vs. Ag/AgCl, KClsat (Co–Cr) and 0.05 V vs. Ag/AgCl,
bined electrode (HI 11310) connected to a pH-meter (type MV 870, KClsat (Ni–Cr), respectively, suggesting the existence of a trans-
Pracitronic, Germany), by addition of small volumes of diluted H3PO4. passivation process associated with the oxygen evolution or dissolu­
The studied values were the following: pH 7 (simulating a normal body tion. In the case of Co–Cr alloy (Fig. 1A), it is well-known that chromium
environment), pH 4.5 (digestion situation), and pH 3 (inflammatory metal in the presence of air, spontaneously forms a protective passive
conditions) [6]. oxide layer of Cr2O3, according to reaction 1 [29,30]:
All experiments were performed at ambient temperature and Cr þ H2O → Cr2O3 þ 6 Hþ þ 6 e (1)
repeated three times, using to each experiment a renewed surface of
metallic alloy electrode. Also, at potentials more positive than 0.8 V vs. Ag/AgCl, KClsat, the
trans-passive oxidation of chromium(III) oxide to form soluble Cr(VI)
species, could occurring following the reaction 2 [29,30]:
2.3. Preparation of alloys electrode
Cr2O3 þ 5 H2O → 2 CrO24 þ 10 Hþ þ 6 e (2)
The dental Ni–Cr and Co–Cr alloys as provided by the producer were
melted and casted as cylinders samples in the same conditions as the In the case of Ni–Cr alloy (Fig. 1B), the trans-passive oxidation of the
dental devices. Than, they were cut into small pieces to a reduced size of Ni oxide film starts at more negative values of potential (i.e., 0.05 V vs.
3 mm diameter and introduced into a Teflon tube, allowing just a Ag/AgCl, KClsat) than in the case of Co–Cr alloy and C1 peak is probably
controlled surface of 0.0765 cm2 to be exposed to the artificial saliva due to the reduction of the previous formed passive film (for example
solution. Before using, all metallic alloys electrodes were mechanically NiO, Ni2O3 or Ni3O4 [30,31]). In the cathodic region a hysteresis of the
cleaned with silicon carbide papers of successively gradations 500, 800, current intensity exists just before the apparition of the C1 peak. The
1000, 2400 and 4000, then well-rinsed with distilled water and dried. reduction process occurs with a smaller intensity of C1 peak at pH 3,
than at pH 7. As seen from Fig. 1A and B for both studied alloys the shape
of voltammogram doesn’t change with the modification of the pH of
2.4. Surface characterization artificial saliva.

The morphology of the alloys surfaces before and after 24 h of im­


mersion in artificial saliva (pH 6) were investigated by scanning electron 3.2. The potential in open circuit (OCP)
microscopy (SEM) using a SU8230 scanning electron microscope
(Hitachi, Japan). Spectra of energy-dispersive X-ray spectroscopy The value of the potential in open circuit (OCP) is an indicator of the
(EDXS) and chemical maps for the elements were acquired using a Dual anticorrosive tendency of metal and has an irreversible character
EDX System (X-Max N100TLE Silicon Drift Detector) from Oxford because the instability of the metal–solution interface. For all alloy
Instruments. samples and all studied pH values, the OCP increased with the time for
stabilization in the first 600 s and remain stable until 3600 s, as in the
3. Results and discussion case of Ni–Cr alloy (Fig. 2). This behavior is explained din literature
[32–34], by the formation of a passive film on the alloy surfaces. The
3.1. Electrochemical behavior of metallic alloys by cyclic voltammetry increase of OCP value is attributed to thickening of the passive film,
which becomes more protective, and so conferring to the alloy more
Fig. 1A–B presents the cyclic voltammograms of Co–Cr and Ni–Cr anticorrosive properties. Also, for both studied alloys, the highest value

392
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Table 1 which gives only information related to the Gibbs free energy of the
Influence of artificial saliva solution’s pH and of stabilization time for OCP on electrochemical system [23,35].
the polarization resistance (Rp) values of Co–Cr and Ni–Cr alloys. Experimental
conditions: see Fig. 3.
3.3. Polarization resistance (Rp)
Type of electrode pH Time for OCP/s Rp/kΩ/cm2 R/no of points

Co–Cr 3 600 13827 0.9962/237 Fig. 3 exhibits the linear polarization resistance (LPR) curves for
3 3600 13105 0.9953/231
Ni–Cr in artificial saliva of pH 3, for different interval of time of stabi­
3 9000 13877 0.9968/232
7 3600 3391 0.9894/132
lization of OCP. As seen in Fig. 3, the OCP values of Ni–Cr alloy shifts to
Ni–Cr 3 600 1331 0.9978/231 more negative values, with the increase of the stabilization time of OCP
3 3600 1256 0.9970/187 (similar behavior for Co–Cr alloy, results not shown).
3 9000 1317 0.9974/235 The polarization resistance (Rp) is calculated as the reverse value of
7 3600 2949 0.9988/237
the slopes of linear I - E plots and the obtained data normalised with the
electrode surface were summarised in Table 1. Comparing the Rp values
of OCP is at pH 7, proving that at pH 3 both studied alloys have the of dental alloys made of precious metals (1500–4000 kΩ/cm2) [3] with
tendency to be oxidized or/and to form soluble compounds. the studied alloys (i.e., 1256 kΩ/cm2 for Ni–Cr and 13105 kΩ/cm2 for
In order to establish if a dental alloy will suffer corrosion under Co–Cr, respectively at pH 3 and time for OCP of 3600 s), it can be
certain condition of the oral cavity, the OCP value determination is concluded that the resistance to corrosion of the Ni–Cr alloys and the
necessary, but not sufficient, because OCP is a thermodynamic measure precious metals are quasi similar. Analysing the data from Table 1, it can
be observed that at pH 7, the Rp value is quasi similar at Ni–Cr and Co–Cr

Fig. 4. Influence of stabilization time for OCP on Tafel plots at Co–Cr (A) and Ni–Cr (B) alloys in artificial saliva solution of pH 3. Influence of pH values on Tafel
plots at Co–Cr and Ni–Cr alloys in artificial saliva solution (C). Experimental conditions: electrolyte, artificial saliva pH 3 (A, B, C) or pH 7 (C, see inset); scan rate,
0.5 mV s 1; interval of potential � 0.200 V vs. OCP; stabilization time for OCP, 3600 s (C).

393
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Table 2
Electrochemical parametres obtained by Tafel extrapolation method from polarization curves of Co–Cr and Ni–Cr alloys. Experimental conditions: see Fig. 4.
Type of electrode pH Time for OCP/s icorr/ βc/ β a/ Rp/ Ecorr/ Corrosion rate/μm/year
μA/cm2 V/dec V/dec kΩ V

Co–Cr 3 600 0.074 0.043 0.041 147.5 0.116 1.102


3600 0.059 0.035 0.040 142.5 0.117 0.880
9000 0.067 0.049 0.036 163.7 0.115 0.992
7 3600 0.057 0.050 0.054 28.1 0.164 0.874
Ni–Cr 3 600 0.202 0.029 0.031 27.2 0.129 3.021
3600 0.063 0.022 0.014 30.5 0.080 0.945
9000 0.154 0.047 0.030 55.4 0.102 2.298
7 3600 0.174 0.051 0.033 57.9 0.059 2.586

alloys. In an acidic environment (pH 3), irrespective the time for stabi­ materials compromises the passive oxide film by breakdown, leading to
lization of OCP, the Rp values are quite different, being 10 times greater a non-uniform corrosion process having a low value of Rp [20]. At pH 3,
in the case of Co–Cr alloy than in the case of Ni–Cr, anticipating a weaker either for Ni–Cr or for Co–Cr, the smallest value of Rp was estimated at a
resistance to corrosion for Ni–Cr alloy. It is well known that the ions time for stabilization of OCP of 3600 s. However, the time for stabili­
released from alloys interact with the ions from the artificial saliva to zation of OCP was established to be 3600s, as a compromise between the
form oxide based passive film, which prevent and reduce the corrosion duration of experiment and the value of Rp [33].
tendency. Also, the variation of pH of saliva in the case of porous

Fig. 5. The Nyquist (A, B) and Bode (C, D) plots of the Co–Cr (A, C) and Ni–Cr (B, D) dental alloys in artificial saliva solution of pH 3 ( , ■, black) and 7 (✯, ★,
red), at initial ( black,✯ red) and after 24 hs (■ black,★ red) of experiment. Experimental conditions: frequencies, 10 kHz - 10 mHz; amplitude, 0.1 V; room
temperature. The line is the fitting curve of the experimental data using the simplified Randles standard equivalent circuit model Rs(Q Rct) (see inset A). (For
interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

394
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Table 3
Fitted EIS parameters of Co–Cr and Cr–Ni alloys using an equivalent circuit model (Rs(Q Rct)). Experimental conditions: see Fig. 5.
pH Time/hs Rs/Ω⋅cm2 Q/S sn/cm2 n Rct/kΩ⋅cm2 Cdl/μF⋅cm 2
χ2
Electrode Co–Cr
3 0 35.46 (�8.14) 3.391 *⋅10 5 (�6.91) 0.807 83.7⋅ (�9.38) 43.5 0.0515
3 10 18.97 (�8.33) 3.30 *⋅10 5 (�7.79) 0.810 138 (�7.66) 47.2 0.0840
3 24 11.32 (�5.90) 3.424 *⋅10 5 (�7.43) 0.806 262.2 (�7.30) 58.1 0.0969
4.5 0 26.54 (�2.96) 8.62 *⋅10 5 (�2.65) 0.790 41.9 (�5.31) 121.2 0.0083
4.5 10 11.32 (�5.44) 6.16 *⋅10 5 (�3.48) 0.812 74.9 (�7.96) 87.8 0.0178
4.5 24 8.25 (�6.15) 7.11 *⋅10 5 (�3.49) 0.792 60.2 (�7.85) 104.1 0.0175
7 0 16.81 (�5.80) 3.64 *⋅10 5 (�4.21) 0.888 160.2 (�9.13) 45.5 0.0297
7 10 9.76 (�7.52) 3.48 *⋅10 5 (�4.46) 0.877 214.3 (�8.69) 46.07 0.0351
7 24 7.285 (�6.74) 3.93⋅*10 5 (�3.52) 0.854 156.5 (�9.26) 53.7 0.0211
Electrode Ni–Cr
3 0 3.70 (�6.20) 162.7 *⋅10 5 (�4.05) 0.557 1.4 (�2.31) 318.5 0.0112
3 10 3.85 (�4.20) 11.66 *⋅10 5 (�2.54) 0.820 26.5 (�4.56) 149.3 0.0085
3 24 3.14 (�4.13) 12.21⋅* 10 5 (�2.27) 0.818 28.9 (�4.42) 161.7 0.0072
4.5 0 17.17 (�1.96) 32.79⋅* 10 5 (�1.91) 0.729 10.3 (�3.86) 515.3 0.0039
4.5 10 12.16 (�2.93) 14.71 * 10 5 (�2.25) 0.755 22.0 (�4.36) 212.1 0.0060
4.5 24 7.89 (�3.11) 14.55⋅* 10 5 (�2.14) 0.771 26.7 (�4.74) 218.1 0.0061
7 0 7.37 (�3.81) 47.72⋅* 10 5 (�3.03) 0.657 4.1 (�4.58) 675.1 0.0073
7 10 9.09 (�4.40) 3.69 * 10 5 (�7.80) 0.912 170.7 (�7.97) 44.0 0.0084
7 24 7.77 (�9.10) 4.82⋅* 10 5 (�7.42) 0.892 121.6 (�7.23) 59.7 0.0864

in brackets: � relative standard error (RSD) expressed in percent (%).

3.4. Polarization curves compact passive layer on the Co–Cr alloy, irrespective of the pH of saliva
solution. In the case of Ni–Cr alloy (time for OCP of 3600 s), the increase
The polarization curves recorded in the range of �200 mV vs. OCP of the pH of the artificial saliva lead to an increase of the corrosion
and plotted in semi-logarithmic graphs (Tafel plots) are presented in current density (from 0.063 μA/cm2 at pH 3 to 0.173 μA/cm2 at pH 7)
Fig. 4 A-C. Using Tafel extrapolation method, corrosion potential (Ecorr, and a similar increase of the corrosion rate (from 0.945 μm/year at pH 3
in Volts), corrosion current intensity (Icorr, in Ampers), and the anodic/ to 2.586 μm/year at pH 7) values.
cathodic Tafel constants (βa, βc in Volts/decade) values could be esti­
mated. Polarization resistance (Rp in ohm) was calculated by
3.5. EIS measurements
Stern–Geary equation (3) [36]:
βa βc Electrochemical impedance spectroscopy (EIS) is a complex tech­
Rp ¼ (3)
2:3 Icorr ðβa þ βc Þ nique used to evaluate the behavior of an electrochemical system, by
modeling the real system in terms of an equivalent circuit. In order to
This equation shows that corrosion current and corrosion resistance
obtain further information on the corrosion resistance properties of the
have inversely proportional values, thus high corrosion current of a
dental alloys, AC impedance measurements were recorded every hour
metal implies that it has low corrosion resistance [13].
within 24 h, at both studied metallic alloys surfaces immersed in arti­
The corrosion rate (μm/year) is obtained from the corrosion current
ficial saliva of pH 3, 4.5 and 7, for a time for stabilization of OCP of
density (icorr in μA/cm2) through Faraday’s law and the density of the
3600 s. Usually a stabilization time of 1 h was used for obtaining the OCP
alloy. Being proportional parameters, either corrosion rate or corrosion
value [37, 33]. Because the surface active surface is very small, a 3600 s
current density can be used to explain the corrosion behavior of the
interval time for obtaining the OCP value is considered to be optimal for
studied alloy samples.
reaching a quasi-equilibrium state at the electrode/solution interface
As seen in Fig. 4 A-C, for both studied alloys, the polarization curves
[33]. Fig. 5 presents the results as Nyquist and Bode diagrams. The
have a similar behavior, irrespective of pH or time for stabilization of
obtained impedance spectrum was fitted using a non-linear regression
OCP. Moreover, on the anodic range a short increase of the current was
method and the values of the parameters of the simplified Randles
observed, suggesting the acceleration of the oxidation process due to a
standard equivalent circuit model (i.e., Rs(Q Rct)) [9,33,38], were
local corrosion phenomenon, especially for the Ni–Cr alloy (irrespective
summarised in Table 3 for the data recorded at 0, 10 and 24 h (inter­
pH or time of OCP), but also for Co–Cr (pH 3, and 3600 s for OCP). At pH
mediary results not shown).
3, for the Co–Cr alloy irrespective of the stabilization time for OCP, the
This wide used equivalent electrical circuit consists of an uncom­
values of the corrosion densities (icorr), corrosion potentials (Ecorr) and
pensated solution resistance (Rs) in series with the parallel combination
corrosion rates have low variation (Table 2). However at an OCP sta­
of the constant phase element (Q) and a charge transfer resistance (Rct).
bilization time of 9000 s, the Ni–Cr alloy shows values for icorr and rate
The Rct is a quantitative parameter, which could be a measure of
of corrosion which are twice than for Co–Cr alloy. The existing differ­
corrosion resistance of passive oxide thin film acting as a barrier to
ences between the values of above mentioned parameters suggest that
electron flow (resistor) between the alloy/electrolyte interface, and
the passive films formed on the Ni–Cr dental material do not have the
provides a value of the degree of protection furnished by the passive
same stability and these materials present differences in corrosion re­
oxide thin film formed on the interface [8,22,33,34,39,40].
sistances. Also, at the shortest stabilization time for OCP of 600 s, the
The use of a constant phase element (Q) instead of an “ideal”
highest value of corrosion current density and of the calculated corro­
capacitor was necessary because of the non-homogeneous character of
sion rates is obtained, irrespective the pH value of artificial saliva or the
the passive oxide thin film present on metallic alloys surface and due to
nature of metallic alloys (Fig. 4C, Table 2), probably due to the fact that
the deviations from 1 observed in Bode plots [8].
the passive protective film is not formed in such short time. Moreover, in
The equivalent capacitance of the Q was modelled as a non-ideal
the case of Co–Cr alloy (time for OCP of 3600 s), the corrosion current
capacitor containing the capacitance of the double layer Cdl and a
densities and the corrosion rates calculated at pH 3 and pH 7 are quasi-
roughness factor n, using the formula Q ¼ (Cdl Rct)1/n/Rct [20,41]. The
similar (i.e., 0.059 and 0.057 μA/cm2 or 0.88 and 0.874 μm/year,
roughness factor n is a number placed in the range 0–1, and it gives
respectively). The behavior could be explained by the existence of a
information about surface roughness (i.e., n ¼ 1 is the case of a perfectly

395
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Fig. 6. SEM images (magnified 100 nm) of Co–Cr (A, C) and Ni–Cr (B, D) fresh polished (A, B) and after corrosion test (C, D) during 24 h in artificial saliva pH 7, and
the corresponding EDXS spectra. Inset: magnified 100 μm.

396
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Table 4
Chemical composition for the alloys given by the manufacturers and obtained from EDXS spectra before and after 24 h corrosion test, in artificial saliva of pH 7.
Alloy Chemical composition, %wt

Co Cr Ni Mo Cu C Mn Si Al O others

Co–Cr given by manufacturer 65 29 – 5 – 0.4 0.25 0.35 – – -


Co–Cr before corrosion test 63 33.1 – 3.5 – – – – – – –
Co–Cr afer corrosion test 46.6 23.8 – 8 – 17.5 – 2 0.8 1.3 –
Ni–Cr given by manufacturer – 25 63 10 – – – 1.5 – – Mn, Al, Zr, Ce, La
Ni–Cr before corrosion test – 18.4 52.5 – 7.6 22.7 21.6 – – – –
Ni–Cr afer corrosion test – 10.3 27.7 – 4.2 – 12.1 1.8 9.4 11.9 –

smooth surface of an ideal capacitor, and n ¼ 0 is the case of an ideal 3.6. Surface morphology and EDXS spectra
resistance, respectively) [42,43].
From the data presented in Table 3 it can be observed that Rs value SEM coupled with EDXS techniques were used for investigation the
doesn’t changed considerably during the experiment, because the fac­ morphologies and the chemical contents of the Co–Cr and Ni–Cr surfaces
tors affecting its value like the stability of exposed surface and ions before (fresh polished) and after the corrosion test of 24 h, in artificial
concentration in artificial saliva solution doesn’t changed. saliva of pH 7.
When a charge transfer reaction occurring to a metallic surface can Analyzing the SEM images for Co–Cr alloy (Fig. 6 A, C), the fresh
be neglected as in the case of corrosion process, a protective film is polished and the corroded surface having a passive oxide layer show
formed on metallic alloy surface, and the corresponding estimated homogenous morphology, irrespective of the magnification level. In the
impedance could be attributed to the impedance of passive layer [39]. case of Ni–Cr, the morphology is quite different from the case of Co–Cr,
In Fig. 5 A-B, the semicircle observed in the Nyquist plot for each the initial surface showing an important porosity consisting in small
metallic alloy surface, was mainly explained to the presence of the craters having holes (Fig. 6 B). The density of craters and the depth of
surface passive layer, which is characterized by a corrosion resistance of holes increase after corrosion test (Fig. 6 D), probably because the
passive layer. As shown the values from Table 3, for both studied accumulation of the ions from artificial saliva at the pores causes the
metallic alloys and irrespective of the pH value of artificial saliva, the Rct passivation film’s breakdown, which in turn accelerates corrosion
value increases after 24 h of contact with the artificial saliva, proving the phenomena.
increase of a the thickness of the passive film on the alloy surface. After The EDS spectra for Co–Cr and Ni–Cr alloys (Fig. 6A–D) show similar
24 h, the alloy most resistive against corrosion, having the greatest Rct chemical composition with those provided by the manufacturer
value [13,18], and consequently, the low rate of released metallic ions (Table 4). The differences between the chemical composition, before
into the body fluid or cell tissue [7,8,17] is the Ni–Cr alloy at pH 7 and after corrosion test, could be due either to the formation of the
(121.6 � 8.79 kΩcm2) and the Co–Cr at pH 3 (262.2 � 19.14 kΩcm2) passive oxide film after interaction between the ions released from alloy
respectively, probably due to the fact that in the specified conditions the and the ions from the artificial saliva solutions during immersion period,
alloys were well protected by the formed passive film. However, at all or to the influence of the pH of solution [30]. The major changes in the
pH (3, 4.5 or 7), the Co–Cr alloy has better corrosion resistance (higher composition of the alloys, especially in the case of Ni–Cr alloy could
Rct) than Ni–Cr based alloys, probably due to the formation of a more dramatically affect the biocompatibility of the employed alloys in the
adherent oxide layer on its surface. Also, for both alloys samples, at all oral cavity.
studied pH, the Rct values increase with the time of study, probably due
to the formation of a Cr-rich passive oxide film, which has a higher value 4. Conclusions
of the resistance to the charge transfer [39]. These values are in accor­
dance with the Rp values obtained by LPR curves. As well-known, corrosion products may affect the biocompatibility
The Chi-square distribution test (χ2) having value between 10 2 - of dental alloys due to wet environment, contact with various types of
10 4, and the relative standard errors (RSD) expressed in percent (%) of ions, pH changes, mechanical or electrochemical wear, which lead to the
all calculated values presented in Table 3 are less than 10%, indicating release of metal ions, local reactions of inflammation of surrounding
excellent agreement between the experimental data and model values, tissue or systemic reactions (e.g., allergic reactions). In this context,
and proving the right choice of using the constant phase element in the Ni–Cr and Co–Cr alloys were exposed to aerated artificial saliva solu­
equivalent circuit model [7,17]. tions of different pHs and the corrosion resistances was investigated by
Table 3 shows that in all pH solutions, Ni–Cr and Co–Cr alloys exhibit electrochemical techniques.
a roughness factor n ¼ 0.56 � 0.89, indicating that the corrosion inter­ The spontaneously formation of a passive oxide layer, was examined
face deviated from a pure ideal capacitor behavior. The phase angle by OCP, linear polarization resistance measurements, Tafel plots and EIS
maximum (θmax) < 90� is probably due to the oxide passive layer [7]. spectra and the obtained results are complementary and in accordance
From the Bode plot (Fig. 5C), for Co–Cr alloy in artificial saliva of with the SEM and EDXS analysis and with the practical life.
different pH, at initial and after 24 h of experiment, it was observed that The EIS experiments, very well fitted using a simplified Randles
the slope of the impedance diagram in the middle frequency range (i.e., standard equivalent circuit model (Rs(Q Rct)), show that at all studied
1–100 Hz) was around 0.80. After 24 h of experiments, the phase angle pH (3, 4.5 or 7), the Co–Cr alloy has better corrosion resistance (Rct)
maxima of the Co–Cr alloy become larger at pH 3 and the values of than Ni–Cr based alloys; the Rct values increasing with the time of
impedance gradually increase [7]. The maximum phase angle is of experiment.
approximately 65� (at pH 3) and 75� (at pH 7), and dropped at about 40� The polarization curves represented by Tafel plots showed that
at lower frequencies. At same pH, irrespective of time of experiment, the irrespective of alimentation habitude of population which can change
behavior is similar [17,22]. Contrarily, the behavior of Ni–Cr alloy is the pH value of saliva, the corrosion rate is under 1 μm/year in the case
similar at the same moment of experiment, irrespective of the pH of of Co–Cr alloy, proving that this alloy is more resistant in time (than the
saliva (Fig. 5D). Therefore, it is evident the passive oxide film of the Ni–Cr alloy) and is recommendable for successful treatment of patients
Co–Cr alloy maintained stable corrosion resistance values in the given with dental prosthesis having metal frameworks.
experimental condition (pH 3), whereas that of the Ni–Cr alloy became
significantly unstable or defective [9].

397
G.L. Turdean et al. Materials Chemistry and Physics 233 (2019) 390–398

Acknowledgments [21] I. Golvano, I. Garcia, A. Conde, W. Tato, A. Aginagalde, J. Mech. Behav. Biomed.
Mater. 49 (2015) 186–196.
[22] X.-Z. Xin, J. Chen, N. Xiang, Y. Gong, B. Wei, Dent. Mater. 30 (2014) 263–270.
The authors acknowledge dr. Lucian Barbu-Tudoran from National [23] M. Talha, Y. Ma, P. Kumar, Y. Lin, A. Singh, Colloids Surfaces B Biointerfaces 176
Institute for Research and Development of Isotopic and Molecular (2019) 494–506.
Technologies (Cluj-Napoca, Romania) for providing SEM and EDXS [24] T. Puskar, D. Jevremovic, R.J. Williams, D. Eggbeer, D. Vukelic, I. Budak, Materials
7 (2014) 6486–6501.
analysis. [25] Y.-Q. Wu, J.A. Arsecularatne, M. Hoffman, J. Mech. Behav. Biomed. Mater. 44
(2015) 23–34.
References [26] S.B. Rao, R. Chowdhary, Int. J. Dentistry (2011), https://doi.org/10.1155/2011/
397029.
[27] G.M.O. de Queiroz, L.F. Silva, J.T. Lima Ferreira, J.A. da Cunha, P. Gomes,
[1] C.P. Bergmann, A. Stumpf, Dental Ceramics, Dental Ceramics. Microstructure, L. Sathler, Braz. Oral Res. 21 (2007) 209–215.
Properties and Degradation, Springer-Verlag Berlin Heidelberg, 2013, p. 9. [28] F. Mansfeld, Corrosion 29 (1973) 397–402.
[2] M. Roach, Dent. Clin. N. Am. 51 (2007) 603–627. [29] N. Kovacevic, B. Pihlar, V.S. Selih, I. Milosev, Acta Chim. Slov. 59 (2012) 144–155.
[3] L. Reclaru, R.E. Unger, C.J. Kirkpatrick, C. Susz, P.-Y. Eschler, M.-H. Zuercher, [30] T. Saleh, J.A. Mars, N. Thovhogi, D. Gihwala, A.A. Baleg, M. Maaza, J. Den. Oral
I. Antoniac, H. Lüthy, Mater. Sci. Eng. C 32 (2012) 1452–1460. Health 1 (2015) 9.
[4] G. Saravanan, S. Mohan, J. Alloy. Comp. 522 (2012) 162–166. [31] M.A. Ameer, E. Khamis, M. Al-Motlaq, Corros. Sci. 46 (2004) 2825–2836.
[5] R.I.M. Asri, W.S.W. Harun, M. Samykano, N.A.C. Lah, S.A.C. Ghani, F. Tarlochan, [32] R.A. Rodríguez-Díaz, A.L. Ramirez-Ledesma, M.A. Aguilar-Mendez, J. Uruchurtu
M.R. Raza, Mater. Sci. Eng. C 77 (2017) 1261–1274. Chavarin, M.A. Hern� andez Gallego, J.A. Juarez-Islas, Int. J. Electrochem. Sci. 10
[6] P. Mo�cnik, T. Kosec, J. Kova�c, M. Bizjak, Mater. Sci. Eng. C 78 (2017) 682–689. (2015) 7212–7226.
[7] R.W.-W. Hsu, C.-C. Yang, C.-A. Huang, Y.-S. Chen, Mater. Chem. Phys. 93 (2005) [33] L. Porojan, C.E. Savencu, L.V. Costea, M.L. Dan, S.D. Porojan, Int. J. Electrochem.
531–538. Sci. 13 (2018) 410–423.
[8] D. Mareci, G. Bolat, R. Chelariu, D. Sutiman, C. Munteanu, Mater. Chem. Phys. 141 [34] C.E. Savencu, L.V. Costea, M.L. Dan, L. Porojan, Int. J. Electrochem. Sci. 13 (2018)
(2013) 362–369. 3588–3600.
[9] C. Qian, X. Wu, F. Zhang, W. Yu, J. Prosthet. Dent 116 (2016) 112–118. [35] R. Galo, L.A. Rocha, A.C. Faria, R.R. Silveira, R. Faria Ribeiro, M.G. Chiarello de
[10] D.C. Hansen, Interface. Summer (2008) 31–33. Mattos, Mater. Sci. Eng. C 45 (2014) 519–523.
[11] *** European Parliament and Council Directive 94/27/EC, Off. J. Eur. [36] Y. Zhang, Z. Xiao, Y. Zhao, Z. Li, Y. Xing, K. Zhou, Mater. Chem. Phys. 199 (2017)
Communities, 1994, pp. 1–2. L188 22. 7, https://eur-lex.europa.eu/legal-content 54–66.
/EN/TXT/PDF/?uri¼CELEX:31994L0027&from¼EN. (Accessed 7 May 2019). [37] D. Mareci, D. Sutiman, A. Cailean, G. Bolat, Bull. Mater. Sci. 33 (2010) 491–500.
[12] *** Directive 2011/24/EU of the European Parliament and of the Council of 9 [38] Y.J. Yang, Z.P. Zhang, Z.S. Jin, W.C. Sun, C.Q. Xia, M.Z. Ma, X.Y. Zhang, G. Li, R.
March 2011 on the application of patients’ rights in cross-border healthcare, Off. J. P. Liu, J. Alloy. Comp. 782 (2019) 927–935.
Eur. Communities (2011) 45–65. L88 4. 4, https://eur-lex.europa.eu/legal-content [39] R. Galo, R.F. Ribeiro, R.C.S. Rodrigues, L.A. Rocha, M.G.C. de Mattos, Braz. Dent. J.
/EN/TXT/PDF/?uri¼CELEX:32011L0024&from¼EN. (Accessed 7 May 2019). 23 (2012) 141–148.
[13] S.H. Tuna, N.O.€ Pekmez, I. Kürkçüoglu, J. Prosthet. Dent 114 (2015) 725–734.
[40] K.J. Qiu, W.J. Lin, F.Y. Zhou, H.Q. Nan, B.L. Wang, L. Li, J.P. Lin, Y.F. Zheng, Y.
[14] A.Y. Musa, M. Behazin, J.C. Wren, Electrochim. Acta 162 (2015) 185–197. H. Liu, Mater. Sci. Eng. C 34 (2014) 474–483.
[15] A.Y. Musa, J.C. Wren, Corros. Sci. 109 (2016) 1–12. [41] Y. Lu, S. Guo, Y. Yang, Y. Liu, Y. Zhou, S. Wu, C. Zhao, J. Lin, J. Alloy. Comp. 730
[16] M. Behazin, J.J. No€ el, J.C. Wren, Electrochim. Acta 134 (2014) 399–410. (2018) 552–562.
[17] M.-H. Hong, T. Hanawa, S.H. Song, B.K. Min, T.-Y. Kwon, J. Mater Res Technol. 8 [42] G.L. Turdean, C.I. Fort, V. Simon, Electrochim. Acta 182 (2015) 707–7144.
(2019) 1587–1592. [43] C.I. Fort, G.L. Turdean, R. Barabas, D. Popa, A. Ispas, M. Constantiniuc, Studia UBB
[18] E. Zhang, C. Liu, Mater. Sci. Eng. C 69 (2016) 134–143. Chem 64 (2019) 125–133.
[19] Y.L. Wang, Q. Wang, H.J. Liu, C.L. Zeng, Corros. Sci. 109 (2016) 43–49.
[20] M. Bahraminasab, M. Bozorg, S. Ghaffari, F. Kavakebian, J. Alloy. Comp. 777
(2019) 34–43.

398

You might also like