Abstract
The amperometric biosensor was created using a flat sheet of the Cu-Zn-Al shape memory alloy, with a shiny surface and uniform thickness below 1 mm. The high biocompatibility and a large specific surface area for enzyme loading are evaluated. In vitro biomedical corrosion testing of samples revealed successful immobilization of catalase, which has undergone quasi-reversible electron transfer from the surface and saline solution. A catalase that had been immobilized retained its basic structure and bioactivity and demonstrated a remarkable electrocatalytic response to hydrogen peroxide reduction. The reduction of hydrogen peroxide on the catalase-modified Cu-25.38Zn-3.3Al alloy was investigated using hydrodynamic amperometry in both the absence and presence of l-cysteine and K2[B3O3F4OH] inhibitors. Catalytic reduction currents have increased as a result of the gradual increase in hydrogen peroxide concentration. The study on enzyme activity inhibition has shown a lower corrosion rate of catalase-modified bioalloy than the initial sample because inhibitor ions occupy all sites of the immobilized enzyme.
1 Introduction
Historically, the development of human society has been linked to the ability of people to produce and process materials to meet their needs. Scientists and engineers have found and designed tens of thousands of materials in the modern era by understanding the interrelationships of the processing, structure, properties and performances. Shape memory alloys are essential materials developed and studied in the last several decades due to their remarkable multifunctional capabilities based on unusual deformation recoverability related to thermomechanical or magnetic stimuli (Chowdhury 2018). The shape changes associated with the martensitic transformation induced by temperature or stress observed following the loading path and the thermomechanical history of the material are one-way shape memory, two-way shape memory, and pseudoelasticity or superelasticity (Dasgupta 2014; Jani et al. 2014). The one-way shape memory effect is characterized by preserving the deformed condition after removing an external force and recovering the original shape upon heating (Jani et al. 2014). Nevertheless, the two-way shape memory phenomenon is associated with shape remembering at high and low temperatures (Jani et al. 2014). Temperature-independent shape memory is known as superelasticity (Wayman and Duerig 1990), referring to the unloading plateau, whereas recovering an apparent plastic deformation just by unloading at a constant temperature, referring to any non-linearity during unloading, is known as pseudoelasticity (Duerig and Zadno 1990).
The developed technologies for manufacturing Ni-Ti, Cu-based, and Fe-based shape memory alloys have enabled the application of those materials in various sectors, from automotive and aerospace to robotics and biomedical science (Aydoğdu et al. 2014; Canbay et al. 2019; Jani et al. 2014). Although different in vitro and in vivo studies have shown good corrosion resistance and biocompatibility of Ni-Ti alloys (Zheng et al. 2011) for designing medical devices such as guide wires, stents, filters, catheters and implants (Colic et al. 2010), Cu-Zn-Al shape memory alloys are attractive for study as their alternative due to lower cost (Olajide et al. 2019). The corrosion behavior of Cu-based shape memory alloys has been evaluated in industrial and intravenous fluids (Alaneme et al. 2013; Olajide et al. 2019), chloride media and borate buffer solutions containing chloride (Alfantazi et al. 2009), Hank’s solution, and Ringer physiological solution (Nady et al. 2016). However, the study of Cu-based alloys as supports material for enzyme immobilization has not gained significant attention, although the electrochemical biosensors containing a biological recognition element are developed and tested for different applications (Ronkainen et al. 2010).
The electrochemical biosensors combine the sensitivity of electrochemical methods with a selectivity of a biological recognition element such as enzymes, proteins, antibodies, nucleic acids, or cells (Ronkainen et al. 2010). The biosensor transforms chemical information into an analytically valuable signal (Thevenot et al. 1999) related to the concentration of the analyte being studied (Ronkainen et al. 2010). It must be particular, selective, accurate, and reproducible with the decreased influence of possible interfering species (Pundir et al. 2018). Since implementing the enzymes in their native form is accompanied by difficulties limiting their application in biochemical, biomedical, biotechnological, and industrial food fields, they are immobilized onto or into a solid matrix (Basak et al. 2013; Cetinus and Oztop 2003). The enzyme immobilization presents the confinement of the enzyme to a matrix/support but is different from that for substrates and products (Datta et al. 2013). Therefore, the most critical factors influencing the biosensor analytical performances are choosing the appropriate matrix and the immobilization procedure (Datta et al. 2013; El Harrad et al. 2018). The research on the electrochemistry of enzymes can provide a good model for investigating enzymatic activity and electron exchange in biological systems (Rusling 1998).
The monofunctional catalases, bifunctional catalases-peroxidases, and manganese catalases are recognized as the classes of enzymes that may be employed in biotechnology as biocatalysts (Grigoras 2017). The heme enzymes are interesting for biotechnological applications due to their diversity of physiological roles, abundance among the naturally occurring cofactors in metalloproteins, and facilitated monitoring, isolation, and characterization (Zuccarello et al. 2021). Catalase (CAT) is a heme-containing protein belonging to the oxidoreductase group and contains four identical subunits equipped with a high-spin Fe(III)-protoporphyrin IX (Murthy et al. 1981; Prakash et al. 2002). It has been found in bacteria, fungi, plants, and mammals, while its activity has been studied in atherosclerosis, neurodegenerative diseases, and cancer cells (Pudlarz et al. 2018). Catalase metabolizes the hydrogen peroxide into oxygen and water without forming free radicals (Murthy et al. 1981). Hydrogen peroxide (H2O2) is a simple compound exhibiting several essential properties such as oxidizing property, gas formation on decomposition, source of free radicals, and effects on biological processes (Pundir et al. 2018). The abnormal expression level of H2O2 in the cells is related to different diseases such as Alzheimer’s disease, cardiovascular disorders and cancer (Liu et al. 2017). It has a crucial role in medical diagnostics, clinical research and environmental applications (Chen et al. 2012), and different industrial sectors, including food processing, paper, textile, pharmaceuticals, cleaning, and disinfection products (Patel et al. 2021).
The development of new nanoscale devices intended for biological, medical and electronic applications depends on combining the biomolecules and new components of biomaterials (Patolsky et al. 2004; Willner 2002). Therefore, we have developed a hypothesis that if catalase was successfully immobilized on the surface of the Cu-Zn-Al shape memory alloy, a reliable biosensor would be developed to determine hydrogen peroxide. Hence, an amperometric biosensor was designed by immobilizing catalase on the sheet surface produced from Cu-25.38Zn-3.3Al alloy using a Nafion matrix. The electrocatalytic reduction of hydrogen peroxide on the catalase-modified Cu-25.38Zn-3.3Al alloy was examined by hydrodynamic amperometry in the absence and the presence of l-cysteine and K2[B3O3F4OH] inhibitors. In vitro testing was performed to evaluate the success of enzyme immobilization on a Cu-25.38Zn-3.3Al alloy using linear polarization and impedance spectroscopy and to determine the kinetic parameters and the type of inhibition carrying out the chronoamperometric technique.
2 Materials and methods
2.1 Materials and chemicals
The shape memory alloy with a composition of Cu-25.38Zn-3.3Al (wt.%) was manufactured by melting high-purity copper, Cu-Zn, and Cu-Al pre-alloys in a graphite crucible in the resistance-heated furnace. A charcoal cover was utilized to prevent copper’s oxygen absorption and melt loss. The molten alloy was poured at 700 °C into preheated (120 °C) graphite mold of dimensions 14 mm × 65 mm × 100 mm. After machining, the cast ingot was cut into cylindrical samples with a 35 mm diameter which were subjected to the homogenization treatment at 850 °C for 2 h, followed by quenching into a water bath at room temperature. The homogenized specimens were hot forged at 830 °C after heating for 1 h, resulting in plates with a thickness of 4.0 mm. The multi-pass cold rolling was accompanied by intermediate annealing treatments at 500 °C for 20 min, followed by air cooling to produce the test specimens with a final thickness of 0.3 mm. Rolled sheets were quenched in a water bath at 17 °C ± 2 °C, after annealing at 890 °C for 10 min.
For designing the biosensor and its testing, the following were used: Enzyme catalase (CAT) (c100-50 MG; Sigma-Aldrich, Buchs, Switzerland); KH2PO4 and Na2HPO4 (Fisher Chemical, Wien, Austria); hydrogen peroxide (H2O2) p.a. 30% (Sigma-Aldrich, Buchs, Switzerland); l-cysteine (Sigma Aldrich, EC 200-158-2); Nafion 5% (Sigma-Aldrich, Buchs, Switzerland). Halogenated boroxine dipotassium-trioxohydroxytetrafluorotriborate K2[B3O3F4OH] as an inorganic boron derivative has been synthesized by the reaction of potassium hydrofluoride (KHF2) with boric acid in a molar ratio of 2:3 at room temperature as described in the literature (Ryss and Slutskaya 1951).
2.2 Characterization techniques
The microstructures of the as-cast and homogenized specimens, as well as rolled and quenched plates, were analyzed with a Zeiss optical microscope. The samples were cut and cold-mounted in epoxy resin, followed by grinding using progressively finer silicon carbide (SiC) grinding papers. Samples were polished and etched using a solution of 5 g FeCl3 + 30 mL HCl + 100 mL H2O, which was used to reveal the microstructure of the cast and homogenized samples. The rolled and quenched specimens were electropolished (voltage: from 20 V to 70 V, solution: 300 mL HNO3 and 600 mL CH3OH, time: 10 s–60 s; cathode of stainless steel) and then etched with a solution of 5 g FeCl3 + 30 mL HCl + 100 mL H2O.
An X-ray diffractometer using Cu-Kα radiation examined the structural properties of quenched samples. The values of the angle φ between lines linking the closest neighbors in the 18R martensite’s basal plane were determined as follows:
where
Using the four-point probe resistance measurement technique, a continuous, direct current of 1.5 A was supplied to the specimen (130 mm × 4 mm × 0.3 mm), and the resulting voltage drop was measured. The martensitic and reverse transformation temperatures were determined based on the change in electrical resistivity with temperature during cooling and heating cycles. Continuous acquisition and data processing were carried out using an amplifier, A/D converter, and relevant software executed on a personal computer. The data processing was accomplished using the original software application. Both applications leveraged the programming and numerical computing platform MATLAB.
At room temperature, mechanical testing of quenched sheets was conducted using an FPZ HECKERT RAUENSTEIN 100/1 standard testing machine until fracture. Five repeat tensile tests were performed to guarantee the reliability of data obtained for the yield stress, tensile strength, and elongation.
Tensile testing was used to investigate shape memory recovery until the tensile strain reached 2.9%, at which point the load was removed. Following unloading, the specimens were heated to 50 °C above Af temperature. Before and after heating the samples, the shape recovery ratio was determined by measuring the distance between two bars within the gauge range.
2.3 Catalase immobilization
An amperometric biosensor for determining the H2O2 was formed by immobilizing catalase on the sheet surface produced from Cu-25.38Zn-3.3Al alloy using a Nafion matrix (Figure 1) as described in the literature (Herenda et al. 2018; Ostojic et al. 2017).

Schematic representation of an amperometric biosensor for the detection of H2O2.
2.4 Electrochemical measurements
The characterization of examined samples was based on the measurements performed on a PAR 263A potentiostat/galvanostat with a conventional three-electrode system. The reference electrode was a saturated Ag/AgCl electrode with a Pt counter, while a sample of Cu-25.38Zn-3.3Al bioalloy was used as a working electrode. The following electrochemical methods were carried out for the measurements: cyclic voltammetry, linear polarization (Tafel), chronoamperometry, and impedance spectroscopy. Cyclic voltammetry was used to examine the immobilization, the effect of different substrate concentrations on enzyme activity, and the thickness of the enzyme film on the electrode surface. All cyclic voltammetry tests were conducted in the phosphate buffer (pH 7) in the potential range of −1.2 V–0.7 V versus Ag/AgCl and at a scan rate of 50 mV/s. In addition, cyclic voltammetry was used to evaluate the effects of different scan rates from 10 mV/s to 90 mV/s. The success of enzyme immobilization on a Cu-25.38Zn-3.3Al bioalloy was examined by linear polarization and impedance spectroscopy. The chronoamperometric technique was implemented for the determination of kinetic parameters, the Michaelis-Menten constant (Km) and the maximum current under saturated substrate conditions (Imax), which is equivalent to the maximum reaction rate (Vmax) and resolve the type of inhibition (Adeyoju 1995). Chronoamperometric measurements were performed in the electrochemical cell containing 25 mL of the phosphate buffer solution at a constant potential of 0.9 V applied to the working electrode and at the constant stirring rate of 400 rpm. The reaction was monitored in different conditions, i.e., in the absence of l-cysteine and boroxine and in different concentrations of l-cysteine and boroxine.
3 Results and discussion
3.1 Optical microscope observation and X-ray diffraction analysis
The microstructures of as-cast and homogenized specimens and rolled and quenched sheets of Cu-25.38Zn-3.3Al bioalloy are shown in Figure 2. The light needles of α-solid solution were observed in the β-phase matrix, while a small number of dendrites were observed in the as-cast samples (Figure 2a). In order to improve the technological plasticity and stability of the mechanical properties as well as reduce the directionality, the bioalloy was homogenized at 850 °C for 2 h and cooled into a water bath at room temperature. A microstructural study of homogenized samples indicated α-phase as the dominating phase and a small amount of martensite (Figure 2b). In terms of future plastic deformation, the homogeneity of the structure is superior to that of the as-cast samples. In light of the fact that one of the needed features of Cu-Zn-Al alloys is a fine grain structure that increases grain boundary strength (Tadaki 1998), cold working was selected as the most effective approach for refining the grain size. The samples were subjected to cold rolling and intermediate annealing treatment at 500 °C for 20 min, followed by air cooling. The intermediate annealing was required after every two or three passes to produce samples with a flat, shiny surface and uniform thickness below 1 mm and avoid crack occurrence and grain coarsening. After annealing, the specimens were cooled in the air to eliminate the internal stresses generated during cold deformation passes. The resulting sheets exhibited a dual-phase microstructure consisting of a minor quantity of β-phase combined with a workable α-phase (Figure 2c). Cold rolling resulted in the elongation of the grains.

Optical micrograph of the (a) as-cast sample, (b) homogenized specimen, (c) final cold-rolled sheet, and (d) quenched sheet of bioalloy Cu-25.38Zn-3.3Al.
After solution treating at 890 °C for 10 min and quenching into a water bath at 17 °C ± 2 °C, blocks of parallel-side martensite plates were formed (Figure 2d). Besides 94.7% of martensite, the quenched sample exhibited approximately 5.3% of the β-phase with DO3 structure (a = 0.5852 nm). The M18 R-type martensite lattice parameters are as follows: a = 0.4464 nm, b = 0.5273 nm, c = 3.7077 nm, and β = 88.35°, while φ = 61.13°. The 2H-type martensite has lattice parameters of a = 0.3166 nm and b = 0.5177 nm.
3.2 The phase transformation temperatures
By monitoring the change of electric resistance with temperature, transformation temperatures of quenched samples were found. The martensitic transition in the rolled and quenched samples began around 257 °C (Ms) and ended at 202 °C (Mf). The quench rate, solution treatment duration, and temperature all substantially impacted the martensitic transformation in the samples studied. The reverse transition of martensite was seen at temperatures ranging from 223 °C (As) to 316 °C (Af temperature).
3.3 Mechanical behavior and shape memory recovery
Figure 3 shows the stress-strain curve of a quenched sample of the bioalloy Cu-25.38Zn-3.3Al. After the initial linear part caused by martensite phase elastic deformation, the stress plateau may be recognized, implying that the martensite reorientation process is at a strain of 0.7%–4.7%. Deformations that occur with subsequent stress application are related to the permanent deformation of the martensite.

Stress-strain curve for Cu-25.38Zn-3.3Al bioalloy.
Tension along the rolling direction till fracture was evaluated on five quenched specimens. The average values of the five measurements were used to calculate the yield stress, tensile strength, and elongation. The yield stress was determined to be 222.0 MPa ± 9.5 MPa, and the ultimate tensile strength to be 481.0 MPa ± 7.8 MPa. The elongation was determined to be 7% ± 1%.
The shape recovery curve shown in Figure 3 indicates a recoverable strain of up to 2.9%, indicating that such an essential thermomechanical response of Cu-25.38Zn-3.3Al shape memory alloy is an unusual deformation associated with the plateau area. Deformation of about 2.9% recovered between 223 °C (As) and 316 °C (Af) temperatures after applied stress has been removed, and samples were heated at 370 °C (temperature which was for 50 °C above Af temperature). Since martensite plates are the only ones that can recover their original configuration by retracing the transformation path during the reverse transformation, as suggested by Ghosh et al. (1986), martensite plates in the microstructure of the quenched sample (Figure 2d) that are oriented favorably concerning the stress axis are primarily responsible for shape recovery.
3.4 Electrochemical characterization
The cyclic voltammograms of a Cu-25.38Zn-3.3Al bioalloy and catalase-modified Cu-25.38Zn-3.3Al bioalloy in phosphate buffer pH 7 at 50 mV/s are shown in Figure 4. The electrochemical response and redox potential formation on the catalase-modified bioalloy may be recognized (Epc = −0.52 V and Eac = 0.05 V). The redox activity depends on the pH and the electrolyte solution composition (Mažeikiene et al. 2003). Different film components causing the potential shifts may interact with the catalase or affect the electrode’s electric double-layer (Salimi et al. 2007).

Cyclic voltammograms of bioalloy samples in phosphate buffer solution pH 7 at a scan rate of 50 mV/s: (a) without immobilized enzyme, (b) with immobilized enzyme.
The effect of the scan rate in the phosphate buffer solution in the absence of oxygen was investigated to obtain the catalase kinetic parameters of on the bioalloy film. Figure 5 shows cyclic voltammograms recorded at various scan rates. A linear relationship (R2 = 0.9698) can be seen in the plot of peak currents versus scan rates (Figure 6), implying that catalase adsorbed on the surface undergoes quasi-reversible electron transfer from the surface of the bioalloy and saline solution. ΔEP was increased with increasing scan rate. The peak-to-peak potential separation values were proportional to the square root of the scan rate. By using the Laviron method for a typical thin-layer electrochemical system (Laviron 1979a,b), the heterogeneous electron transfer rate constant (ks) of the catalase immobilized on the electrode (bioalloy) was estimated. The transfer coefficient and heterogeneous electron transfer rate constant of catalase were about 0.45 and 3.7 ± 0.1 s−1, respectively. Results indicate that the transfer of catalase electrons to the bioalloy is an easy and uninterrupted process.

Cyclic voltammograms of bioalloy in the phosphate buffer at different scan rates.

The plot of anodic peak currents versus square root of scan rate.
The cyclic voltammograms for the reduction of hydrogen peroxide on the catalase-modified bioalloy electrode at different concentration ranges are shown in Figure 7. The proportionality of the catalytic peak currents with the concentration of hydrogen peroxide was found with a linear range from 0.3 mM to 3 mM (Figure 8), for which the linear regression equation may be written as follows:

Cyclic voltammograms of immobilized bioalloy in the presence of different substrate concentrations of H2O2 at a scan rate of 50 mV/s.

The catalytic currents response versus hydrogen peroxide concentrations.
Cyclic voltammetry response for all, including higher hydrogen peroxide concentration, showed a leveling-off tendency of the typical Michaelis-Manten process of formation of the enzyme-substrate complex (Stryer 1988).
In order to confirm the success of the immobilization of catalase on the bioalloy, we have also used the impedance spectroscopy technique and linear polarization resistance measurements. The oxidation-reduction processes were best observed on the bioalloy at the potential of 0.9 V during our preliminary research. Therefore, electrochemical impedance diagrams were recorded at a constant voltage of 0.9 V. A frequency ranged from 10,000 Hz–0.01 Hz with a 5 mV amplitude sine wave. The Nyquist diagram in Figure 9 represents the obtained results of the electrochemical impedance spectroscopy of bioalloy in the absence and presence of the immobilized enzyme. The initial parts of the curves are not clearly visible and experimentally recorded, which indicates that they correspond to very high frequencies (above 10,000 Hz). Also, it can be noticed that the experimental curves do not form closed semicircles, which indicates that the missing parts of the curves are related to frequencies lower than 0.01 Hz. The obtained impedance spectrum revealed a successful enzyme immobilization onto the Cu-Zn-Al alloy sheet. In order to confirm the inhibitory activity of analyzed substances, the impedance spectra have been recorded under the same conditions. The measurement was performed in the absence of K2[B3O3F4OH] as well as its presence at a concentration of 0.028 mM for a fixed concentration of immobilized catalase of 7.01 × 10−3 g/mL, with the addition of a substrate at a concentration of 3.3 mM. The values obtained by the intersection of semicircles with the real axis indicate the electrolyte resistance Rel. In Figure 10, for measurement without K2[B3O3F4OH], the electrolyte resistance Rel value was found as 135 Ω, while 130 Ω was determined for measurement with K2[B3O3F4OH]. A difference between the initial measurements can be seen better in Figure 11, which presents the enlarged view of the initial part of the Nyquist diagram presented in Figure 10. The charge transfer resistance value through the electrolyte Rct, for measurement in the absence of K2[B3O3F4OH], was recorded as ∼2200 Ω. After adding K2[B3O3F4OH], the charge transfer resistance value was determined as ∼2000 Ω. Obtained results indicate the inhibitory effect of K2[B3O3F4OH] on catalase activity. Since the corrosion rate of the material in saline can be compared to the polarization resistance, it may be seen that the corrosion rate of Cu-25.38Zn-3.3Al bioalloy is higher than that of catalase-modified Cu-25.38Zn-3.3Al bioalloy in the presence of inhibitors.

Nyquist diagram of the impedance spectrum of bioalloy in the (a) absence and (b) presence of immobilised enzyme at a potential of 0.9 V.
![Figure 10:
Nyquist diagram of the impedance spectrum of catalase activity without () and in the presence of K2[B3O3F4OH] 0.028 mM ().](/document/doi/10.1515/corrrev-2022-0025/asset/graphic/j_corrrev-2022-0025_fig_026.jpg)
Nyquist diagram of the impedance spectrum of catalase activity without () and in the presence of K2[B3O3F4OH] 0.028 mM (
).
![Figure 11:
Enlarged view of the initial part of the Nyquist diagram of the impedance spectrum of catalase activity without () and in the presence of K2[B3O3F4OH] in the concentration of 0.028 mM ().](/document/doi/10.1515/corrrev-2022-0025/asset/graphic/j_corrrev-2022-0025_fig_027.jpg)
Enlarged view of the initial part of the Nyquist diagram of the impedance spectrum of catalase activity without () and in the presence of K2[B3O3F4OH] in the concentration of 0.028 mM (
).
Obtained charge transfer resistance value through the electrolyte Rct in this work is almost twice as lower as the values determined by the researchers who have investigated the corrosion and corrosion inhibition of Mg/Mn alloy in 3.5% NaCl solutions by 5-(3-aminophenyl)-tetrazole (APT) after different exposure intervals (Babouri et al. 2017; El-Sherif and Almajid 2011). The observed difference is attributed to the influence of chloride ions. The increase in resistance of Mg with APT concentration is a consequence of the adsorption of its molecules on the alloy surface due to the chloride ions’ attack. On the other hand, the semicircles at high frequencies are generally associated with the relaxation of electrical double-layer capacitors, while the diameters of the high-frequency semicircles can be considered as the charge transfer resistance (El-Sherif and Almajid 2011).
The success of enzyme immobilization on the bioalloy at a scan rate of 0.166 mV is illustrated in Figure 12. The value of the corrosion potential EP was −0.03 V compared to the alloy EP (0.07 V). The drop in potential indicates that the biosensor is cross-linked in the polymer, as seen by the curves in Figure 12. Electron mobility and diffusion are impeded, resulting in a decrease in potential. The cathodic polarization of Cu-25.38Zn-3.3Al bioalloy was better than for immobilized alloy from the polarization curves. The passivation of Al was visible. Since the surface ionization of Al was more substantial than that of Zn in Cu-25.38Zn-3.3Al bioalloy samples, it could be concluded that this compact Al layer inhibited cathodic oxygen depletion. The corrosion rate of catalase-modified bioalloy was lower than that of Cu-25.38Zn-3.3Al bioalloy due to the lower potential. The corrosion potentials of the material tested in the same medium could be a symbol of thermodynamic stability. In the physiological saline solution, the thermodynamic stability of the Cu-25.38Zn-3.3Al bioalloy was more significant than the catalase-modified bioalloy due to the possibility of uneven surface immobilization.

Tafel polarisation curves of bioalloy in the presence of biosensor CAT (b) and absence of it (a).
The trend of activation-passivation appeared in the anodic polarization of Cu-25.38Zn-3.3Al bioalloy, while the trend of passivation did not occur in the catalase-modified bioalloy. In Cu–Zn–Al shape memory alloys, elements Zn and Al have stronger metal activity than Cu and may be easier to ionize in solutions (Chen et al. 2005). Adding Zn and Al as alloying elements to copper to form Cu-25.38Zn-3.3Al alloy contributes to decreasing the tendency of anodic polarization and improving the performance of cathodic polarization, which was observed during examining pure copper and the 70Cu-26Zn-4Al shape memory alloy in the simulated uterine fluid by Chen et al. (2005). Immobilization success and biocorrosion parameters indicate that the direct electrochemistry of enzymes can provide a good model for mechanistic studies of their electron transfer activity in biological systems, as discussed by Mani et al. (2012). The CAT enzyme immobilized in the film on the surface of Cu-25.38Zn-3.3Al alloy shows well-defined and direct electrochemical reactions. It retains its primary structure, bioactivity, and good stability.
Cu-based alloys are prone to aging treatment which affects shape memory behavior. The properties of these alloys are improved by the presented method of immobilization and reaction of alloy ions and iron metal ions in the enzyme. Therefore, it is essential to emphasize that the analysis of the corrosion potential is critical before shape memory alloys are placed in biomedical or industrial applications. It was found that alloys with shape memory have higher corrosion resistance than traditional ones due to the elastic behavior of the polycrystalline structure (Al-Hassani et al. 2017; Raheem et al. 2010). Raheem et al. (2010) observed that austenite’s current density (633.62 μA/cm2) was greater than that of martensite (336.45 μA/cm2) in Cu-Al-Ni shape memory alloy.
As one of the most applied biosensor techniques (Amine and Mohammadi 2019; Zhang et al. 2008), hydrodynamic amperometry has been used to study the electrocatalytic reduction of hydrogen peroxide on the CAT film of the bioalloy. The constant potential of the modified electrode (mixing at 400 rpm) at 0.9 V versus Ag/AgCl was applied. The catalytic reduction current was monitored while the different hydrogen peroxide aliquots were added. The increase in catalytic reduction currents was caused by a gradual increase in the concentration of hydrogen peroxide from 6 mM to 32 mM in a physiological saline solution. From the Lineweaver-Burk diagram presented in Figures 13 and 14, the inhibition type has been defined and values of maximum current (Imax) and Michaelis-Menten constant (Km) were calculated in the absence as well as the presence of l-cysteine and K2[B3O3F4OH] inhibitors.

Lineweaver-Burk diagram for the determination of kinetic constants (Imax and Km) in the absence and presence of different concentrations of l-cysteine.
![Figure 14:
Lineweaver-Burk diagram for the determination of kinetic constants (Imax and Km) in the absence and presence of different concentrations of K2[B3O3F4OH].](/document/doi/10.1515/corrrev-2022-0025/asset/graphic/j_corrrev-2022-0025_fig_030.jpg)
Lineweaver-Burk diagram for the determination of kinetic constants (Imax and Km) in the absence and presence of different concentrations of K2[B3O3F4OH].
In the enzyme reactions carried out, the competitive type of inhibition is recognized from the Lineweaver-Burke diagram shown in Figure 13. The obtained Imax value is constant (0.8 μA) while the Km values were increased with raising the inhibitor concentration (0.0184 mM; 0.0396 mM; 0.0935 mM), indicating that the inhibitor binds to the active site of the enzyme on the bioalloy. The ESI complex can be formed by binding l-cysteine to the previously formed enzyme-substrate complex. The reaction proceeds in the direction that the ESI complex yields the product but less than the reaction without an inhibitor. With the addition of K2[B3O3F4OH] to the reaction mixture, the inhibitor has bound competitively to catalase, as shown in Figure 14. The calculated values of the kinetic constants Imax and Km are changed by adding different concentrations of inhibitors. While Imax values were 3.3 μA; 1.4 μA and 0.66 μA, the values of Km were 0.016 mM; 0.0075 mM and 0.0043 mM. The initial formation of the ES binding and inhibition complexes is required for this modality of inhibitors (Copeland 2013). Therefore, after the initial non-covalent binding of the substrate to the enzyme’s active site, these inhibitors affect the catalysis steps. The binding of boroxine to the ES complex is occurred according to the mechanism (Marangoni 2003) as follows:

(3)
where E, enzyme; S, substrate; I, inhibitor; ES, enzyme–substrate complex; ESI, ES–inhibitor; P, product.
Based on the mechanism and the obtained Km constant, our research proved that K2[B3O3F4OH] has higher catalytic stability due to lower Km values, which is in good agreement with the theory. Todorovic et al. (2014) evaluated the biocompatibility of Cu-Al-Ni shape memory alloys in vitro and proved that those alloys almost entirely reduced the metabolic activity of peripheral blood mononuclear cells (PBC).
4 Conclusions
In this paper, in vitro biomedical corrosion behavior and inhibition of enzyme activity were studied to develop the new biosensor applying Cu-25.38Zn-3.3Al shape memory alloy. High-quality sheets with flat, shiny surfaces and uniform thickness below 1 mm were produced by multi-pass cold rolling of hot forged samples without cracks and coarse grains. After quenching into a water bath at room temperature, a dominantly martensitic structure exhibited a yield stress of 222.0 MPa ± 9.5 MPa, and tensile strength of 481.0 MPa ± 7.8 MPa, while the elongation of 7% ± 1%, and recovery strain of up to 2.9% were evaluated.
An amperometric biosensor was designed by immobilizing catalase on the sheet surface using a Nafion matrix. The catalase adsorbed on the surface of the alloy underwent quasi-reversible electron transfer from that surface and saline solution. The results indicate that immobilized catalase’s primary structure and biological activity were retained, showing a great electrocatalytic response to hydrogen peroxide reduction.
The hydrodynamic amperometry was used to study the electrocatalytic reduction of hydrogen peroxide on the catalase-modified Cu-25.38Zn-3.3Al alloy in the presence of l-cysteine and K2[B3O3F4OH] inhibitors, as well as without inhibitors. In the enzyme reactions carried out, the competitive type of inhibition occurred. In the presence of inhibitors, catalase-modified alloy’s corrosion rate was lower than ordinary Cu-25.38Zn-3.3Al alloy because inhibitor ions occupy all sites of immobilized enzyme on the alloy. The kinetic parameters of Km and Imax inhibition show that the inhibited catalase-modified alloy follows the Michaelis-Menten model. On the basis of the mechanism and the obtained Km constant, it was proved that K2[B3O3F4OH] has higher catalytic stability than l-cysteine.
Based on the performed study, it could be concluded that Cu-based shape memory alloys with medical applications can be used to immobilize. In order to develop an efficient carrier for enzyme immobilization for final application in the food processing and pharmaceutical industry, future efforts would be towards designing the chemical compositions and thermomechanical procedures for producing biocompatible and nontoxic surface of Cu-based shape memory alloys.
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: None declared.
-
Conflicts of interest: The authors declare no conflicts of interest regarding this article.
References
Adeyoju, O.O. (1995). Development of horseradish peroxidase and tyrosinase-based organic-phase biosensors, Doctoral dissertation. Dublin City University.Search in Google Scholar
Alaneme, K.K., Sulaimon, A.A., and Olubambi, P.A. (2013). Mechanical and corrosion behaviour of iron modified Cu-Zn-Al alloys. Acta Metall. Slovaca 19: 292–301, https://doi.org/10.12776/ams.v19i4.184.Search in Google Scholar
Alfantazi, A.M., Ahmed, T.M., and Tromans, D. (2009). Corrosion behavior of copper alloys in chloride media. Mater. Des. 30: 2425–2430, https://doi.org/10.1016/j.matdes.2008.10.015.Search in Google Scholar
Al-Hassani, E.S., Ali, A.H., and Hatem, S.T. (2017). Investigation of corrosion behavior for copper-based shape memory alloys in different media. Eng. Techn. J. 35: 578–586, https://doi.org/10.30684/etj.35.6a.4.Search in Google Scholar
Amine, A. and Mohammadi, H. (2019). Amperomtry. In: Worsfold, P., Townshend, A., Poole, C., and Miro, M. (Eds.). Encyclopedia of analytical science, 3rd ed Academic Press, Oxford, pp. 85–98.Search in Google Scholar
Aydoğdu, Y., Kürüm, F., Kὂk, M., Yakinci, Z.D., and Aydoğdu, A. (2014). Thermal properties, microstructure and microhardness of Cu-Al-Co shape memory alloy system. Trans. Indian Inst. Met. 67: 595–600, https://doi.org/10.1007/s12666-013-0376-1.Search in Google Scholar
Babouri, L., Belmokre, K., Brioua, S., and Bardeau, J.F. (2017). Study of the corrosion-erosion behavior of Cu70-Ni30 alloy inflowing water containing marble particles. J. Chem. Pharmaceut. Res. 9: 141–152.Search in Google Scholar
Başak, E., Aydemir, T., Dinçer, A., and Becerik, S.Ç. (2013). Comperative study of catalase immobilization on chitosan, magnetic chitosan and chitosan-clay composite beads. Artif. Cells, Nanomed. Biotechnol. 41: 408–413, https://doi.org/10.3109/10731199.2013.796312.Search in Google Scholar PubMed
Canbay, C.A., Karaduman, O., Unlu, N., Baiz, S.A., and Ozkul, I. (2019). Heat treatment and quenching media effects on the thermodynamical, thermoelastical and structural characteristics of a new Cu-based quaternary shape memory alloy. Compos. B Eng. 174: 106940, https://doi.org/10.1016/j.compositesb.2019.106940.Search in Google Scholar
Cetinus, S.A. and Oztop, H.N. (2003). Immobilization of catalase into chemically crosslinked chitosan beads. Enzym. Microb. Technol. 32: 889–894, https://doi.org/10.1016/s0141-0229(03)00065-6.Search in Google Scholar
Chen, B., Liang, C., Fu, D., and Ren, D. (2005). Corrosion behavior of Cu and the Cu–Zn–Al shape memory alloy in simulated uterine fluid. Contraception 72: 221–224, https://doi.org/10.1016/j.contraception.2005.04.006.Search in Google Scholar PubMed
Chen, W., Cai, S., Ren, Q.-Q., Wen, W., and Zhao, Y.-D. (2012). Recent advances in electrochemical sensing for hydrogen peroxide: a review. Analyst 137: 49–58, https://doi.org/10.1039/c1an15738h.Search in Google Scholar PubMed
Chowdhury, P. (2018). Frontiers of theoretical research on shape memory alloys: a general overview. Shape Memory Superelasticity 4: 26–40, https://doi.org/10.1007/s40830-018-0161-4.Search in Google Scholar
Colic, M., Rudolf, R., Stamenkovic, D., Anzel, I., Vucevic, D., Jenko, M., Lazic, V., and Lojen, G. (2010). Relationship between microstructure, cytotoxicity and corrosion properties of a Cu–Al–Ni shape memory alloy. Acta Biomater. 6: 308–317, https://doi.org/10.1016/j.actbio.2009.06.027.Search in Google Scholar PubMed
Copeland, R.A. (2013). Evaluation of enzyme inhibitors in drug discovery: a guide for medicinal chemists and pharmacologists, 2nd ed. John Wiley & Sons, Hoboken, New Jersey.10.1002/9781118540398Search in Google Scholar
Dasgupta, R. (2014). A look into Cu-based shape memory alloys: present scenario and future prospects. J. Mater. Res. 29: 1681–1698, https://doi.org/10.1557/jmr.2014.189.Search in Google Scholar
Datta, S., Christena, L.R., and Rajaram, Y.R.S. (2013). Enzyme immobilization: and overview on techniques and support materials. Biotech 3: 1–9, https://doi.org/10.1007/s13205-012-0071-7.Search in Google Scholar PubMed PubMed Central
Duerig, T.W. and Zadno, R. (1990). An engineer’s perspective of pseudoelasticity. In: Duerig, T.W., Melton, K.N., Stockel, D., and Wayman, C.M. (Eds.), Engineering aspects of shape memory alloys. Butterworth-Heinemann, London, pp. 369–370.10.1016/B978-0-7506-1009-4.50036-6Search in Google Scholar
El Harrad, L., Bourais, I., Mohammadi, H., and Amine, A. (2018). Recent advances in electrochemical biosensors based on enzyme inhibition for clinical and pharmaceutical applications. Sensors 18: 164, https://doi.org/10.3390/s18010164.Search in Google Scholar PubMed PubMed Central
El-Sherif, S.M. and Almajid, A.A. (2011). Corrosion of magnesium/manganese alloy in chloride solutions and its inhibition by 5-(3-aminophenyl)-tetrazole. Int. J. Electrochem. Sci. 6: 2131–2148.10.1016/S1452-3981(23)18172-7Search in Google Scholar
Ghosh, B., Banerjee, M.K., and Seal, A.K. (1986). Shape memory in some copper alloys. Mater. Sci. Technol. 2: 496–499, https://doi.org/10.1179/mst.1986.2.5.496.Search in Google Scholar
Grigoras, A.G. (2017). Catalase immobilization – a review. Biochem. Eng. J. 117: 1–20, https://doi.org/10.1016/j.bej.2016.10.021.Search in Google Scholar
Herenda, S., Ostojić, J., Hasković, E., Hasković, D., Miloš, M., and Galić, B. (2018). Electrochemical investigation of the influence of K2[B3O3F4OH] on the activity of immobilized superoxide dismutase. Int. J. Electrochem. Sci. 13: 3279–3287, https://doi.org/10.20964/2018.04.35.Search in Google Scholar
Jani, J.M., Leary, M., Subic, A., and Gibson, M.A. (2014). A review of shape memory alloy research, applications and opportunities. Mater. Des. 56: 1078–1113, https://doi.org/10.1016/j.matdes.2013.11.084.Search in Google Scholar
Laviron, E. (1979a). The use of linear potential sweep voltammetry and of AC voltammetry for the study of the surface electrochemical reaction of strongly adsorbed systems and of redox modified electrodes. J. Electroanal. Chem. 100: 263–270, https://doi.org/10.1016/s0022-0728(79)80167-9.Search in Google Scholar
Laviron, E. (1979b). General expression of the linear potential sweep voltammograms in the case of diffusion less electrochemical systems. J. Electroanal. Chem. 101: 19–28, https://doi.org/10.1016/s0022-0728(79)80075-3.Search in Google Scholar
Liu, H., Weng, L., and Yang, C. (2017). A review on nanomaterial-based electrochemical sensors for H2O2, H2S and NO inside cells on revealed by cells. Microchim. Acta 184: 1267–1283, https://doi.org/10.1007/s00604-017-2179-2.Search in Google Scholar
Mani, V., Devadas, B., Chen, S.M., and Li, Y. (2012). Immobilization of enzymes and redox proteins and their electrochemical biosensor applications. ECT Trans. 50: 35–41, https://doi.org/10.1149/05012.0035ecst.Search in Google Scholar
Marangoni, A.G. (2003). Enzyme kinetics: a modern approach. John Wiley & Sons, Hoboken, NJ.10.1002/0471267295Search in Google Scholar
Mažeikiene, R., Niaura, G., and Malinauskas, A. (2003). Voltammetric study of the redox processes of self-doped sulfonated polyaniline. Synth. Met. 139: 89–94, https://doi.org/10.1016/s0379-6779(03)00037-7.Search in Google Scholar
Murthy, M.R., Reid, T.J., Sicignano, A., Tanaka, N., and Rossmann, M.G. (1981). Structure of beef live catalase. J. Mol. Biol. 152: 465–499, https://doi.org/10.1016/0022-2836(81)90254-0.Search in Google Scholar PubMed
Nady, H., El-Rabiei, M.M., and Abd El-Hafez, G.M. (2016). Electrochemical stability of Cu-10Ni-10Zn ternary alloys in simulated physiological solutions. J. Bio- Tribo-Corros. 2: 1–12.10.1007/s40735-016-0058-8Search in Google Scholar
Olajide, J.L., Zannu, F.J., Daramola, O.O., Ogunbadejo, A.S., Okotete, E.A., Sadiku, E.R., and Alaneme, K.K. (2019). Morphological characterization, in vitro biomedical corrosion and corrosion behaviour of As-cast Cu-Zn-Al-FeMn alloys in selected intravenous and industrial fluids. Mater. Res. Express 6: 1–9, https://doi.org/10.1088/2053-1591/ab309d.Search in Google Scholar
Ostojic, J., Herenda, S., Bešić, Z., Milos, M., and Galic, B. (2017). Advantages of an electrochemical method compared to the spectrophotometric kinetic study of peroxidase inhibition by boroxine derivative. Molecules 22: 1120, https://doi.org/10.3390/molecules22071120.Search in Google Scholar PubMed PubMed Central
Patel, V., Kruse, P., and Selvaganapathy, P.R. (2021). Solid state sensors for hydrogen peroxide detection. Biosensors 11: 1–31, https://doi.org/10.3390/bios11010009.Search in Google Scholar PubMed PubMed Central
Patolsky, F., Weizmann, Y., and Willner, I. (2004). Long-range electrical contacting of redox enzymes by SWCNT connectors. Angew. Chem., Int. Ed. 43: 2113–2117, https://doi.org/10.1002/anie.200353275.Search in Google Scholar PubMed
Prakash, K., Prajapati, S., Ahmad, A., Jain, S.K., and Bhakuni, V. (2002). Unique oligomeric intermediates of bovine liver catalase. Protein Sci. 11: 46–57, https://doi.org/10.1110/ps.ps.20102.Search in Google Scholar
Pudlarz, A.M., Czechowska, E., Ranoszek-Soliwoda, K., Tomaszewska, E., Celichowski, G., Grobelny, J., and Szemraj, J. (2018). Immobilization of recombinant human catalase on gold and silver nanoparticles. Appl. Biochem. Biotechnol. 185: 717–735, https://doi.org/10.1007/s12010-017-2682-2.Search in Google Scholar PubMed
Pundir, C.S., Deswal, R., and Narwal, V. (2018). Quantitative analysis of hydrogen peroxide with special emphasis on biosensors. Bioproc. Biosyst. Eng. 41: 313–329, https://doi.org/10.1007/s00449-017-1878-8.Search in Google Scholar PubMed
Raheem, K.A., Abid, A., and Al-Tai, T.Z. (2010). The effect of iron addition on the dry sliding wear and corrosion behavior of Cu Al Ni shape memory alloy. Eng. Technol. J. 28: 6888–6902.10.30684/etj.28.24.5Search in Google Scholar
Ronkainen, N.J., Halsall, H.B., and Heineman, W.R. (2010). Electrochemical biosensors. Chem. Soc. Rev. 39: 1747–1763, https://doi.org/10.1039/b714449k.Search in Google Scholar PubMed
Rusling, J.F. (1998). Enzyme bioelectrochemistry in cast biomembrane-like films. Acc. Chem. Res. 31: 363–369, https://doi.org/10.1021/ar970254y.Search in Google Scholar
Ryss, I. and Slutskaya, M. (1951). Report on the platinum sector. Akad. Nauk. SSSR 26: 216–218.Search in Google Scholar
Salimi, A., Sharifi, E., Noorbakhsh, A., and Soltanian, S. (2007). Direct electrochemistry and electrocatalytic activity of catalase immobilized onto electrodeposited nano-scale islands of nickel oxide. Biophys. Chem. 125: 540–548, https://doi.org/10.1016/j.bpc.2006.11.004.Search in Google Scholar PubMed
Stryer, L. (1988). Biochemistry, 3rd ed.. W. H. Freeman, New York.Search in Google Scholar
Tadaki, T. (1998). Cu-based shape memory alloys. In: Otsuka, K. and Wayman, C.M. (Eds.), Shape memory materials. Cambridge University Press, Cambridge, p. 143.Search in Google Scholar
Thevenot, D., Toth, K., Durst, R., and Wilson, G. (1999). Electrochemical biosensors: recommended definitions and classification. Pure Appl. Chem. 71: 2333–2348, https://doi.org/10.1351/pac199971122333.Search in Google Scholar
Todorović, A., Rudolf, R., Romčević, N., Đorđević, I., Milošević, N., Trifković, B., Veselinović, V., and Čolić, M. (2014). Biocompatibility evaluation of Cu-Al-Ni shape memory alloys. Contemp. Mater. 2: 228–238, https://doi.org/10.7251/comen1402228t.Search in Google Scholar
Wayman, C.M. and Duerig, T.W. (1990). An introduction to martensite and shape memory. In: Duerig, T.W., Melton, K.N., Stockel, D., and Wayman, C.M. (Eds.), Engineering aspects of shape memory alloys. Butterworth-Heinemann, London, p. 14.10.1016/B978-0-7506-1009-4.50005-6Search in Google Scholar
Willner, I. (2002). Biomaterials for sensors, fuel cells, and circuits. Science 98: 2407–2408, https://doi.org/10.1126/science.298.5602.2407.Search in Google Scholar PubMed
Zhang, L., Tian, D.-B., and Zhu, J.J. (2008). Direct electrochemistry and electrochemical catalysis of myoglobin – TiO2 coated multiwalled carbon nanotubes modified electrode. Biochemistry 74: 157–163, https://doi.org/10.1016/j.bioelechem.2008.07.003.Search in Google Scholar PubMed
Zheng, Y.F., Zhang, B.B., Wang, B.L., Wang, Y.B., Li, L., Yang, Q.B., and Cui, L.S. (2011). Introduction of antibacterial function into biomedical TiNi shape memory alloy by the addition of element Ag. Acta Biomater. 7: 2758–2767, https://doi.org/10.1016/j.actbio.2011.02.010.Search in Google Scholar PubMed
Zuccarello, L., Barbosa, C., Todorovic, S., and Silveira, C.M. (2021). Electrocatalysis by heme enzymes – applications in biosensing. Catalysts 11: 218, https://doi.org/10.3390/catal11020218.Search in Google Scholar
© 2023 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- Reviews
- A comprehensive review on synergistic and individual effects of erosion–corrosion in ferrous piping materials
- Application of machine learning in material corrosion research
- Corrosion inhibition efficiency and quantum chemical studies of some organic compounds: theoretical evaluation
- Original Articles
- In vitro biomedical corrosion and enzyme activity inhibition on modified Cu-Zn-Al bioalloy
- Imidazoline behavior as corrosion inhibitor in the electrochemical characterization of SCC behavior of an API X70 steel exposed to brine solution
- In-situ visualization of hydrogen atom distribution at micro-indentation in a carbon steel by scanning Kelvin probe force microscopy
- Influence of exposure in a corrosive environment on ultimate stress of heat-treated welded joints of Al–Mg–Si–Сu alloy
- Surface preparation and double layer effect for silane application on electrogalvanized steel
Articles in the same Issue
- Frontmatter
- Reviews
- A comprehensive review on synergistic and individual effects of erosion–corrosion in ferrous piping materials
- Application of machine learning in material corrosion research
- Corrosion inhibition efficiency and quantum chemical studies of some organic compounds: theoretical evaluation
- Original Articles
- In vitro biomedical corrosion and enzyme activity inhibition on modified Cu-Zn-Al bioalloy
- Imidazoline behavior as corrosion inhibitor in the electrochemical characterization of SCC behavior of an API X70 steel exposed to brine solution
- In-situ visualization of hydrogen atom distribution at micro-indentation in a carbon steel by scanning Kelvin probe force microscopy
- Influence of exposure in a corrosive environment on ultimate stress of heat-treated welded joints of Al–Mg–Si–Сu alloy
- Surface preparation and double layer effect for silane application on electrogalvanized steel