Open access peer-reviewed chapter

Exploring Musa paradisiaca Peel Extract as a Green Corrosion Inhibitor for Mild Steel Using Factorial Design Method

Written By

Olusola S. Amodu, Moradeyo O. Odunlami, Joseph T. Akintola, Seteno K. Ntwampe and Seide M. Akoro

Submitted: 25 July 2018 Reviewed: 20 November 2018 Published: 12 February 2019

DOI: 10.5772/intechopen.82617

From the Edited Volume

Corrosion Inhibitors

Edited by Ambrish Singh

Chapter metrics overview

1,701 Chapter Downloads

View Full Metrics

Abstract

The suitability ofMusa paradisiaca (banana) peel extract as a green corrosion inhibitor for mild steel in acidic medium (1 M HCl) was investigated using factorial method of the design of experiment. The effects of two independent variables (concentration of banana peel extract and temperature) on the corrosion inhibition efficiency were investigated. The physicochemical properties of the extract such as surface tension, viscosity, flash point, and specific gravity were determined using standardized methods provided by the American System of Testing Materials (D-971). The relationship between the independent variables and the inhibitor efficiency was modeled by gasometric and thermometric methods. The statistical analysis of the inhibition efficiency was carried out using the “Fit Regression Model” of Minitab® 17.0, while the fitness of the models was assessed by the coefficient of determination (R2) and the analysis of variance (ANOVA). From the results obtained, gasometric method achieved a maximum inhibition efficiency of 66.83%, with an R2 of 90.76%, whereas thermometric method gave a maximum inhibition efficiency of 65.70%, with an R2 of 95.56%. This study shows that banana peel extract has the capacity to prevent the corrosion of mild steel in acidic medium.

Keywords

  • banana peel extract
  • biomass
  • corrosion
  • inhibitors
  • factorial design

1. Introduction

In 2016, NACE International estimated the global cost of corrosion at US$ 2.5 trillion annually. This accounts for about 3.4% of the global gross domestic product (GDP). In the same study, it was discovered that if corrosion prevention best practices are implemented, there could be global savings of between 15 and 35% of the cost of damage [1]. In spite of the technological advancement of this generation, high profile cases of corrosion have continued to emerge [2, 3, 4]. Moreover, the extensive application of acid solutions in industrial cleaning and descaling of mild steel makes metal dissolution a common phenomenon Gadiyar et al. [5]. In order to prolong the lifespan of mild steel, to enhance its viability, and to reduce the high cost of production, practical steps need to be employed in corrosion prevention. Failure to prevent or manage corrosion can result to metal losses, loss of production time, leaking vessels, and unwarranted cleanup costs.

Corrosion is an electrochemical process; it is the propensity for metals to revert to their natural ore state. It takes place in the presence of moisture and oxygen, involving chemical reaction and the flow of electrons on the surface of the corroded cells, which greatly accelerate the transformation of metal back to the low-grade ore. The process involves the oxidation of a metal atom, whereby it loses one or more electrons. The resultant effect of corrosion is metal degradation, that is, the breakup of bulk metal, causing it to lose its useful properties [6]. This electrochemical process, often referred to as galvanic cell, occurs when two different metals in physical or electrical contact are immersed in a common electrolyte with different concentrations. Consequently, the more active metal (anode) gets corroded while the more noble metal (cathode) is protected [7, 8]. The fundamental chemical reactions occurring at the anode and cathode are:

Anode surface : F e F e 2 + + 2 e
Cathode surface : 1 2 O 2 + H 2 O + 2 e 2 OH
At anode : F e 2 + + 2 OH F e O . H 2 O hydrated iron oxide brown rust

Galvanic corrosion is the most common type of corrosion and it occurs regularly in marine vessels, metal structures, and oil pipelines. Furthermore, the phenomenon is commonly observed in water treatment plant, boilers, storage vessels, oil pipelines, etc. In fact, what makes corrosion challenging is that it starts on the internal part of the metal structure, which makes early detection difficult. Other than water and oxygen, galvanic corrosion is affected by: types of metal, agitation, the presence and type of inhibitors, and environmental factors (pH, temperature, humidity, salinity, etc.) [9]. In addition, the dissolution of mild steel in HCl is given as:

2 F e + H + F e 2 + + H 2

The rate of this reaction is dependent on: metal (its position in the electromotive series), acidity, ferrous ion concentration (by the law of mass action, the increase in ferrous ions should correspond to the decrease in rate of corrosion), and hydrogen gas evolution.

Three techniques are often used for the assessment of corrosion rates namely, weight loss technique, electrochemical impedance spectroscopy, and hydrogen gas evolution method. The weight loss method is considered the most fundamental, against which the accuracy of the other methods is determined. However, the limitations of this method are: (1) the weight loss expressed is the average of the weight of the corroding specimen over a period of time but the changes in corrosion rate over this period is not accounted for; (2) in order to accurately determine the weight loss caused by corrosion, all the corroded particles need to be removed from the specimen surface without removing the uncorroded metal, which practically is unrealistic. Electrochemical technique has been successfully used in many corrosion studies to determine the rate of corrosion [4, 7, 10]. Particularly, it has certain advantages over the weight loss method, due to its ease of corrosion rates determination. With this method, instantaneous corrosion rates as well as changes in corrosion rates over a period of time can be determined. However, during the electrochemical dissolution process for some metals and metal alloys, the atypical polarization performance at the anode is a challenge. Moreover, hydrogen removal can occur in two ways: hydrogen gas evolution and depolarization via oxidation by dissolved oxygen or by some other oxidizing agent [11]. Hydrogen gas evolution method seems to be the most significant and reliable in assessing the rate of corrosion as the mole of metal dissolved directly correlates to the amount of hydrogen gas given off [12].

The application of inhibitors is one of the practical ways to protect metals against corrosion, especially in acidic media. Basically, they function by serving as integuments on metal surface thereby preventing it from chloride ions and oxygen dissolution. Corrosion inhibitors find application in minimizing metallic waste in engineering materials, in addition to the advantages of versatility and cost effectiveness when compared to other corrosion protection methods [10, 13]. Most of the effective inhibitors are either from biomass precursors or chemical compounds containing hetero-atoms such as oxygen, nitrogen, and sulfur with multiple bonds in their molecules through which they are adsorbed on the metal surface [14, 15, 16]. This adsorption depends on certain physiochemical properties of the inhibitors: functional group, electron density at the donor atom, n-orbital character, and the electron structure of the molecule. They contain electronegative functional groups and π-electrons in their double bonds, which facilitate their adsorption onto the metal surface. Hence, the strength of an inhibitor to either prevent corrosion reaction from being initiated or slow down the rate of corrosion, is dependent upon the molecular structure of the inhibitor molecules.

The major concern with most chemical inhibitors is their toxicity to the environment. Although many of these synthetic compounds have shown good anti-corrosive activity; their applications have been limited due to environmental considerations [17, 18, 19, 20]. Particularly, inorganic corrosion inhibitors such as lead and chromium have been found to constitute significant health challenge to human when released into the environment. This has necessitated the quest for environmentally benign precursors as corrosion inhibitors, for which a plethora of organic materials have been reported. Some of these materials are plant extracts of kola nut, tobacco, Rosmarinus officinalis, Cassia auriculata, Argemone Mexicana, unripe fruit of Musa acuminate, roasted coffee seed (Coffea arabica), Carica papaya leaves, cannabis plant, pomegranate, etc. [14, 19, 21, 22, 23]. Furthermore, research has revealed that the basic components of plant extracts are alkaloids, steroids, sugars, gallic acid, tannic acid, and flavonoids, which have been found to improve the protection of metal surface against corrosion [24, 25, 26].

The use of natural organic plant extracts as corrosion inhibitors is sometimes referred to as green corrosion prevention. These natural organic compounds are relatively cheaper, nontoxic, and readily available, either as agro-waste and/or agro-industrial waste [27]. Therefore, the aim of this study was to investigate the effectiveness of Musa paradisiaca (banana) peel extract as a green corrosion inhibitor for mild steel in acidic medium. The rate of corrosion was assessed using the hydrogen gas evolution method. In addition, the interaction effects of concentration and temperature variation on the corrosion inhibition were assessed using full factorial design and analyzed with relevant statistical tools.

Advertisement

2. Materials and methods

2.1. Preparation of samples and corrosion test solutions

Banana peels were sourced locally. The peels were washed under running water and air dried until a constant mass was recorded. It was milled into powder using a hammer-mill and ball-mill to achieve finesse of powder (about 0.25 mm diameter size). The powdered peel was extracted using 95% ethanol. Five grams of powder was dissolved in 200 ml ethanol for 14 days and thereafter filtered. The filtrate was rotatory evaporated in order to remove excess ethanol, and then diluted with 1 M HCl in distilled water to obtain the corrosion inhibition test solutions in the concentration ratio of 1.0, 2.5, 5, 7.5, and 10% (v/v). Figure 1 shows the various stages involved in the extraction of banana peels.

Figure 1.

Flow process for banana peel extract (BPE) as corrosion inhibitor.

In addition, the mild steel used was mechanically press-cut into coupons of dimensions 4 × 2.5 × 0.1 cm. Each coupon was degreased by washing with ethanol, dried in acetone, and immediately transferred into the simulated test solutions. Note that the dried coupons can be preserved in a desiccator until use. Similarly, control experiments were set up but without the addition of the inhibitor. All reagents used were of analytical grade. Banana peel is composed of starch (3%), total dietary fiber (43.2–49.7%), crude fat (3.8–11%), crude protein (6–9%), polyunsaturated fatty acids, pectin, micronutrients (K, P, Ca, and Mg), and amino acid [28]. Also, the mild steel sheet used has the following compositions (% wt): Fe—99.3, Mn—0.34, Cu—0.069, Co—0.069, Ca—0.087, Ni—0.043 and Al—0.03.

2.2. Evaluating the physical properties of banana peel extract as a corrosion inhibitor

2.2.1. Viscosity measurement

This was determined by the Cannon-Fenske viscometer and a circulatory bath with temperature control. Viscosity was calculated using ASTM Method D445–97 [29]. The viscosity η of each sample was calculated using the formula below:

η = k ρ T , E1

where k is the instrument constant, ρ is the density of banana peel extract sample, and T efflux time (sec) for banana peel extract sample.

2.2.2. Specific gravity determination

The extract of banana peel was transferred into a narrow glass cylinder (SP0121-V Osaka, Japan) and a hydrometer was set into the sample and allowed to stabilize. The value of the specific gravity was taken from the markings on the stem of the hydrometer at the surface of the extract sample.

2.2.3. Surface tension determination

This study employed the American System of Testing Materials D-971 [29] method. Two grams of banana peel extract was added to 50 ml of distilled water in a 100 ml beaker. A platinum ring was then lowered into the solution of banana peel extract in the beaker. It was then brought up to the water sample interface, where the actual measurement takes place. The force required to pull the ring through the interface was measured by a tension meter as the surface tension of the extract solution (dynes cm−1).

2.2.4. Flash point measurement

The measurement of the flash point for the BPE sample was done using ASTM D-92 method [29]. An open cup containing BPE sample was heated at a specific rate while flame was periodically passed over its surface. The lowest temperature at which the BPE vapor ignites without sustaining the flame was recorded as the flash point.

2.3. Evaluating the corrosion inhibition efficiency of banana peel extract

2.3.1. Gasometric method

This method was adopted in this study as described by Ekpe et al. [23], and carried out at the following temperatures: 303, 308, 313, 318, and 323 K, which were achieved using a water bath. The coupons immersed in the prepared test solutions were recovered after 6 h, washed in detergent solution, and rinsed with distilled water, and air dried. The volume of gas evolved from the cathodic reaction during the corrosion process was determined. Hence, gasometric method correlates the quantity of gas evolved to the rate of corrosion. The graph of the volume of gas liberated per minute gives the rate of gas evolution, while the inhibition efficiency (𝓔) and degree of surface coverage ( θ ) were determined from Eqs. 2 and 3, respectively.

E = 1 V H V 0 H × 100 E2
θ = 1 V H V 0 H E3

where V*H is volume of hydrogen gas evolved at time t in the presence of inhibitor and V0H is the volume of hydrogen evolved in the absence of inhibitor.

2.3.2. Thermometric method

Temperature determination was carried out as reported by Ebenso et al. [30]. Using the value for the rise in temperature per minute, the reaction number (RN) was calculated as shown in Eq. 4:

RN ° C / min = Tm Ti / t E4

where Tm and Ti are the maximum and initial temperatures, respectively, attained by the system and t is the time. Similarly, the inhibition efficiency was determined by the reaction number correlation (Eq. 5).

E = R N o R N i R N o × 100 E5

where RNo is the reaction number of solution without inhibitor, while RNi is the reaction number of solution with inhibitor.

2.4. Effects of concentration and temperature variations on corrosion inhibition

The data obtained from the banana peel extraction experiments were analyzed statistically using factorial method to obtain a linear fit for the inhibition of mild steel corrosion under varied concentration and temperature. The mathematical model was generated by a MINITAB 17.0. Regression analysis was performed to correlate the response variable to the independent variables. The quality of the fit of the model was evaluated using analysis of variance (ANOVA), where the response was the inhibition efficiency; while concentration and temperature of BPE solution were the input variables.

Advertisement

3. Results and discussion

3.1. Physical properties of BPE

The physical properties determined for the banana peel extract are presented in Table 1.

Density (g/L cm3) Dynamic viscosity (cp) Surface tension (dynes cm−1) Specific gravity Flash point (°C)
1.56 32.04 18.0 1.55 237

Table 1.

Physical properties of banana peel extract (BPE).

The viscosity of the banana peel extract (BPE) compares favorably well with those previously reported for BPE and other biomaterials. For instance, similar viscosities were reported for extracts from Citrullus lanatus, Phyllanthus, and banana peels [31, 32]. Viscosity is the property of a fluid that makes it resist flow and sustain frictional force. It is an important property of a good inhibitor, which makes it stick to the surface of metals thereby forming a protective barrier against corrosion. It represents the ability of the extract to adhere longer to metal surfaces, thus enhancing corrosion inhibition. Corrosion inhibitors often contain one or more surfactants, which lowers the surface tension of corrosive fluids [33]. Similarly, the flash point of BPE extract is within the range reported by Kliskic et al. [34] and Betiku et al. [35]. This indicates the flammability or combustibility of the inhibitor.

3.2. Evaluation of BPE efficiency as a corrosion inhibitor

3.2.1. Gasometric method

The efficiency of corrosion inhibition by BPE was determined using gasometric method according to Eq. 1, for the various concentrations of 1.0, 2.5, 5.0, 7.5, and 10.0 g/L (Table 2).

Levels 1 2 3 4 5
Concentration (g/L), X1 1.0 2.5 5.0 7.5 10.0
Temperature (K), X2 303 308 313 318 323

Table 2.

Experimental design.

The volume of gas evolved from the cathodic reaction during the corrosion study (Figure 2) is correlated to inhibition efficiency by gasometric method.

Figure 2.

Volumetric rate of hydrogen gas during corrosion of mild steel in 1 M HCl acid solution.

As shown in Figure 2, the volume of H2 evolved decreased with increasing concentration of the extract. This can be attributed to increased adsorption forces at higher concentrations. However, hydrogen gas evolution increased with time, until after 8 min before equilibration. This is expected since adsorption decreases with time. From the chemical equation, the dissolution of 2 M of ferrous generates one molecule of hydrogen gas. This means that the evolution of one molecule of hydrogen gas corresponds to the dissolution of 2 M of iron. Theoretically, by mole ratio, it implies that the rate of iron dissolution doubles the rate of hydrogen gas evolution. From the anodic and cathodic reactions, the flow of chloride ion from cathode caused the dissolution of iron at the anode and the metal gets corroded, which is the formation of hydrated iron (II) chloride. Hence, prevention of corrosion is possible if the chloride ion is prevented from having contact with the metal surface. The corrosion of mild steel in HCl solutions has been reported to be a first order reaction [36], which increases with increased acid concentration. In addition, the dissolution of iron steel in hydrochloric acid is dependent upon chloride ion over acidic range of pH. Also, the pH of HCl solution decreased with increased immersion time of the mild steel (Figure 3). This was obviously due to increased acidity of the solution.

Figure 3.

Variation of pH of HCl solution with immersion time of mild steel in the presence of banana peel extract (BPE).

Moreover, as described in Section 2.3, the volume of hydrogen gas evolution was used to determine the corrosion inhibition efficiency at different temperatures and concentrations of the biomass extract. Consequently, experimental design was used to assess the effects of these independent parameters that ultimately led to peak process performance and the discovery of optimum conditions. The experimental design was generated using a MINITAB 17.0 software (Stat-Ease Inc., USA). Each variable was analyzed at five levels with a total of 25 experiments being performed representing a full factorial. The system response, which is the corrosion inhibition efficiency, was determined by gasometric and thermometric methods (Tables 3 and 4).

Run order Concentration (g/L) (X1) Temperature (K) (X2) Inhibition, Ɛ (%)
1 7.5 303 64.14
2 5.0 313 51.11
3 1.0 303 44.40
4 10.0 313 61.75
5 2.5 308 49.53
6 7.5 318 49.25
7 7.5 313 53.14
8 2.5 318 40.28
9 5.0 308 56.31
10 2.5 313 44.43
11 10.0 323 49.21
12 1.0 323 24.75
13 2.5 303 54.73
14 10.0 303 72.03
15 1.0 313 34.00
16 7.5 308 59.00
17 7.5 323 44.17
18 1.0 318 29.15
19 5.0 323 41.62
20 5.0 303 61.51
21 1.0 308 39.20
22 5.0 318 29.15
23 10.0 308 66.83
24 10.0 318 54.62
25 2.5 323 34.92

Table 3.

Corrosion inhibition efficiency determination using gasometric method.

Run order Concentration (g/L) (X1) Temperature (K) (X2) Inhibition, Ɛ (%)
1 7.5 303 62.4
2 5.0 313 41.93
3 1.0 303 41.58
4 10.0 313 59.85
5 2.5 308 49.25
6 7.5 318 44.98
7 7.5 313 51.73
8 2.5 318 33.94
9 5.0 308 54.13
10 2.5 313 40.31
11 10.0 323 48.77
12 1.0 323 21.75
13 2.5 303 51.75
14 10.0 303 71.56
15 1.0 313 29.95
16 7.5 308 57.92
17 7.5 323 40.86
18 1.0 318 24.86
19 5.0 323 37.72
20 5.0 303 60.01
21 1.0 308 35.32
22 5.0 318 41.93
23 10.0 308 65.70
24 10.0 318 53.62
25 2.5 323 31.56

Table 4.

Corrosion inhibition efficiency determination using thermometric method.

3.2.2. Thermometric method

Similarly, using thermometric method given in Eq. 4, the percentage corrosion inhibition was evaluated for the concentrations 1.0, 2.5, 5.0, 7.5, and 10.0 g/L, and at various temperatures. The experimental results presented in Tables 3 and 4 show that at constant concentration of the extract; say, at 10 or 7.5 g/L, corrosion inhibition efficiency decreased with increasing temperature. In the same vein, it was observed that at constant temperature; say, 303 K, corrosion inhibition efficiency increased with increasing inhibitor concentration. In a similar study, Gunavathy and Murugavel [37] have reported that the inhibition efficiency of mild steel corrosion in acid medium by Musa acuminata fruit peel extract increases with the increase in concentration but decreases with increase in temperature. The same trend was reported by Mayanglambam et al. [38] for Musa Paradisiaca extract on mild steel in sulfuric acid solution. Lai et al. [39] have also worked on the inhibition of mild steel in HCl acid solution with a synthetic inhibitor, and/or synthetic mixed-type inhibitors as revealed by potentiodynamic polarization measurement. It was equally reported that the efficiency of the inhibitors decreased with increasing temperature as well as acid concentration.

3.3. Modeling and statistical analysis

Suitable statistical models were chosen to model the interactions between the different experimental variables and their effect on the efficiency of corrosion inhibition, based on the “Fit Regression Model” of MINITAB 17.0 (Pen, USA). The response was modeled with a response surface quadratic model and further analyzed by analysis of variance (ANOVA) to assess the significance of each variable on corrosion inhibition. An empirical model that could relate the response measured to the independent variables was obtained using multiple regression analysis. The response (Y), can be represented by the following quadratic model:

Y = α o + i = 1 n α i X i + i = 1 n α ii X 2 i + i = 1 n 1 j = i + 1 n α ij X i X j + ε E6

where X1, X2, X3, …, Xn are the independent coded variables, α0 is the offset term, and αi,αii, and αij account for the linear, squared, and interaction effects, respectively, and ε is the random error. A model reduction may be expedient, if there are many redundant model terms [40].

The statistical model summary based on the Lack-of-Fit Test explained the fitness of quadratic models. Using ANOVA to assess the significance of each variable in the model, empirical quadratic models were obtained from Eq. 6. These models, Eqs. (7) and (8), for gasometric and thermometric methods, respectively, were used to predict the efficiencies of the corrosion inhibition at the various values of the independent variables.

Y = 376.0 + 2.660 X 1 1.0910 X 2 . E7
Y = 376.4 + 2.928 X 1 1.1038 X 2 . E8

The ANOVA of the quadratic regression model for the corrosion inhibition showed the significant level of the model at 90.72 and 95.56% for gasometric and thermometric methods, respectively (Tables 5 and 6). It indicates how well the model fits the experimental data, implying that the total variance in the response could be explained using this model. The closeness in the values of R-sq (adj) and R-sq (pred) in both methods also shows the significance of the model. Regression results often show the statistical correlation and importance between the predictor and response. The coefficient of determination (R2) is the percentage of inhibition efficiency (%) variation that is explained by its relationship with concentration (g/L) and temperature (K). Therefore, the adjusted R2 is the percentage of inhibition efficiency (%) variation that is explained by its relationship with concentration (g/L) and temperature (K), adjusted for the number of predictors in the model. This adjustment is important because R2 for this model increases when a new independent variable is added. The adjusted R2 is a useful tool for comparing the explanatory power of models with different numbers of predictors. P-value for each coefficient tests the null hypothesis that the coefficient has no effect [40].

Source DF Adj SS Adj MS F-value P-value
Regression 2 3372.9 1686.45 107.59 0.000
Conc. of BPE (g/L) 1 1885.6 1885.59 120.30 0.000
Temperature (K) 1 1487.3 1487.31 94.89 0.000
Error 22 344.8 15.67
Total 24 3717.7
Model summary
S R-sq R-sq (adj) R-sq (pred)
3.95907 90.72% 89.88%
Coefficient
Terms Coef SE coef T-value P-value VIF
Constant 376.0 35.1 10.72 0.000
Conc. of BPE (g/L) 2.660 0.243 10.97 0.000 1.00
Temperature (K) −1.091 0.112 −9.74 0.000 1.00

Table 5.

ANOVA for corrosion inhibition efficiency of BPE (gasometric method).

Source DF Adj SS Adj MS F-value P-value
Regression 2 3807.7 1903.87 236.50 0.000
Conc. of BPE (g/L) 1 2284.8 2284.78 283.81 0.000
Temperature (K) 1 1523.0 1522.97 189.18 0.000
Error 22 177.1 8.05
Total 24 3984.9
Model summary
S R-sq R-sq (adj) R-sq (pred)
2.83731 95.56% 95.15% 94.36%
Coefficient
Terms Coef SE coef T-value P-value VIF
Constant 376.0 25.1 14.97 0.000
Conc. of BPE (g/L) 2.928 0.174 16.85 0.000 1.00
Temperature (K) −1.1038 0.0803 −13.75 0.000 1.00

Table 6.

ANOVA for corrosion inhibition efficiency of BPE (thermometric method).

3.4. Graphical representation of the model

Graphical representation allows easy interpretation of experimental results and the prediction of optimal conditions. From the contour plots, Figures 4 and 5, BPE was most efficient at temperatures below 307.5 K and concentrations above 8 g/L. The least inhibitory effect was observed at temperature of 318 K and at lower concentrations. This corroborated the experimental results where the highest corrosion inhibition was 72.03 and 71.56%, for gasometric and thermometric methods, respectively. This peak performance occurred when concentration was 10 mol/L and temperature 303 K. Other higher values determined for inhibition efficiency occurred in the neighborhood of 303 and 308 K, and 10 mol/L. Furthermore, the interactive effect of the concentration and temperature on the system’s response (inhibition efficiency) was assessed by plotting three-dimensional curves of the response against the independent variables (Figures 6 and 7). The response distribution in this experiment with respect to the variation of the independent variables shows that temperature has a greater effect.

Figure 4.

Contour plot showing the effects of concentration and temperature on the efficiency of corrosion inhibition of BPE (Gasometric method).

Figure 5.

Contour plot showing the effects of concentration and temperature on the efficiency of corrosion inhibition of BPE (Thermometric method).

Figure 6.

Surface plot of inhibition efficiency (%) against concentration (g/L) and temperature (K) for gasometric method.

Figure 7.

Surface plot of inhibition efficiency (%) against concentration (g/L) and temperature (K) for thermometric method.

Similar corrosion inhibition efficiencies have been reported for biomass extracts [37, 41]. In some other studies, inhibition efficiencies in the range of 80–90 have been reported for mild steel in HCl solution [42, 43, 44, 45]. In a study carried out by Ong and Karim [46], where the extract of red onion was used to inhibit corrosion of mild steel in HCl solution, an inhibition efficiency of 90% was reported also at temperature of 303 K. Since a combination of factors such as temperature, concentration of inhibitors, and immersion time affects inhibition efficiency, it is pretty difficult to compare extracts of different biomass. However, reports have shown that temperature and concentration of inhibitors are the predominant factors [47].

Advertisement

4. Conclusion

Statistical analysis using full factorial and the Regression Fit Model of MINITAB 17.0 was carried out to assess the effectiveness of Musa paradisiaca (banana) peel extract as a green corrosion inhibitor for mild steel in acidic medium. The effect of concentration of inhibitor and reaction temperature was investigated while the efficiency of corrosion inhibition was evaluated by gasometric and thermometric methods. The system’s response (inhibition efficiency) showed a stochastic distribution with respect to the independent variables, with the highest corrosion inhibition efficiency being 72.03 and 71.56%, for gasometric and thermometric methods, respectively. This peak performance occurred when the concentration was 10 mol/L and temperature 303 K. Furthermore, the ANOVA of the quadratic regression model for the corrosion inhibition showed the significant level of the model at 90.72 and 95.56% for gasometric and thermometric methods, respectively. The response surface as well as the contour plots indicated the extract from the agro-waste was most efficient at temperature below 307.5 K and at concentrations between 8 and 10 g/L. The least inhibitory effect was observed at temperatures above 318 K and at concentrations below 6 mol/L. Banana peel extract is one of those plant extracts that have shown to be promising in green corrosion inhibition.

References

  1. 1. NACE International. International Measures of Prevention, Application and Economics of Corrosion Technology (IMPACT) Study. 2016. Available from: http://inspectioneering.com/ [Accessed: May 23, 2018]
  2. 2. Liu Z, Kleiner Y. State of the art review of inspection technologies for condition assessment of water pipes. Measurement. 2013;46(1):1-15
  3. 3. Kishawy HA, Gabbar HA. Review of pipeline integrity management practices. International Journal of Pressure Vessels and Piping. 2010;87(7):373-380
  4. 4. Iribarren JI, Liesa F, Alemán C, Armelin E. Corrosion rate evaluation by gravimetric and electrochemical techniques applied to the metallic reinforcing structures of a historic building. Journal of Cultural Heritage. 2017;27:153-163
  5. 5. Gadiyar HS, Chintamani D, Gaonkar KB. Chemical cleaning, decontamination and corrosion. Bombay, India: Bhabha Atomic Research Centre; 1991
  6. 6. Rebak RB, Perez TE. Effect of Carbon Dioxide and Hydrogen Sulfide on the Localized Corrosion of Carbon Steels and Corrosion Resistant Alloys. Paper Presented at the CORROSION 2017; 2017
  7. 7. Snihirova D, Taryba M, Lamaka SV, Montemor MF. Corrosion inhibition synergies on a model Al-Cu-Mg sample studied by localized scanning electrochemical techniques. Corrosion Science. 2016;112:408-417
  8. 8. Kim BH, Lim SS, Daud WRW, Gadd GM, Chang IS. The biocathode of microbial electrochemical systems and microbially-influenced corrosion. Bioresource Technology. 2015;190:395-401
  9. 9. El Haleem SA, El Wanees SA, Bahgat A. Environmental factors affecting the corrosion behaviour of reinforcing steel. VI. Benzotriazole and its derivatives as corrosion inhibitors of steel. Corrosion Science. 2014;87:321-333
  10. 10. Tan Y-J, Bailey S, Kinsella B. An investigation of the formation and destruction of corrosion inhibitor films using electrochemical impedance spectroscopy (EIS). Corrosion Science. 1996;38(9):1545-1561
  11. 11. Taneja JN. Dissolution of iron in hydrochloric acid [Electronic Theses and Dissertations]. Canada: University of Windsor; 1967
  12. 12. Song G, Atrens A, StJohn D. An hydrogen evolution method for the estimation of the corrosion rate of magnesium alloys. Magnesium Technology. 2001;2001:254-262
  13. 13. Volk C, Dundore E, Schiermann J, Lechevallier M. Practical evaluation of iron corrosion control in a drinking water distribution system. Water Research. 2000;34(6):1967-1974
  14. 14. Verma C, Olasunkanmi LO, Ebenso EE, Quraishi MA, Obot IB. Adsorption behavior of glucosamine-based, pyrimidine-fused heterocycles as green corrosion inhibitors for mild steel: Experimental and theoretical studies. The Journal of Physical Chemistry C. 2016;120(21):11598-11611
  15. 15. Zheng X, Zhang S, Gong M, Li W. Experimental and theoretical study on the corrosion inhibition of mild steel by 1-octyl-3-methylimidazolium L-prolinate in sulfuric acid solution. Industrial & Engineering Chemistry Research. 2014;53(42):16349-16358
  16. 16. Verma C, Quraishi M, Kluza K, Makowska-Janusik M, Olasunkanmi LO, Ebenso EE. Corrosion inhibition of mild steel in 1M HCl by D-glucose derivatives of dihydropyrido [2, 3-d: 6, 5-d′] dipyrimidine-2, 4, 6, 8 (1H, 3H, 5H, 7H)-tetraone. Scientific Reports. 2017;7:44432
  17. 17. Bethencourt M, Botana F, Calvino J, Marcos M, Rodriguez-Chacon M. Lanthanide compounds as environmentally-friendly corrosion inhibitors of aluminium alloys: A review. Corrosion Science. 1998;40(11):1803-1819
  18. 18. Morad M. Effect of amino acids containing sulfur on the corrosion of mild steel in phosphoric acid solutions containing Cl−, F− and Fe 3+ ions: Behavior under polarization conditions. Journal of Applied Electrochemistry. 2005;35(9):889-895
  19. 19. Abiola OK, Oforka N, Ebenso E, Nwinuka N. Eco-friendly corrosion inhibitors: The inhibitive action of Delonix Regia extract for the corrosion of aluminium in acidic media. Anti-Corrosion Methods and Materials. 2007;54(4):219-224
  20. 20. Nnaji NJ, Ujam OT, Ibisi NE, Ani JU, Onuegbu TO, Olasunkanmi LO, et al. Morpholine and piperazine based carboxamide derivatives as corrosion inhibitors of mild steel in HCl medium. Journal of Molecular Liquids. 2017;230:652-661
  21. 21. Odewunmi N, Umoren S, Gasem Z. Watermelon waste products as green corrosion inhibitors for mild steel in HCl solution. Journal of Environmental Chemical Engineering. 2015;3(1):286-296
  22. 22. Ituen E, James A, Akaranta O, Sun S. Eco-friendly corrosion inhibitor from Pennisetum purpureum biomass and synergistic intensifiers for mild steel. Chinese Journal of Chemical Engineering. 2016;24(10):1442-1447
  23. 23. Ekpe U, Ibok U, Ita B, Offiong O, Ebenso E. Inhibitory action of methyl and phenyl thiosemicarbazone derivatives on the corrosion of mild steel in hydrochloric acid. Materials Chemistry and Physics. 1995;40(2):87-93
  24. 24. Gerengi H, Sahin HI. Schinopsis lorentzii extract as a green corrosion inhibitor for low carbon steel in 1 M HCl solution. Industrial & Engineering Chemistry Research. 2011;51(2):780-787
  25. 25. Muthumanickam S, Jeyaprabha B, Karthik R, Elangovan A, Prakash P. Adsorption and corrosion inhibiting behavior of Passiflora foetida leaf extract on mild steel corrosion. International Journal of Corrosion and Scale Inhibition. 2015;4(4):365-381
  26. 26. Alibakhshi E, Ramezanzadeh M, Bahlakeh G, Ramezanzadeh B, Mahdavian M, Motamedi M. Glycyrrhiza glabra leaves extract as a green corrosion inhibitor for mild steel in 1 M hydrochloric acid solution: Experimental, molecular dynamics, Monte Carlo and quantum mechanics study. Journal of Molecular Liquids. 2018;255:185-198
  27. 27. Amodu OS, Ntwampe SK, Ojumu TV. Emulsification of hydrocarbons by biosurfactant: Exclusive use of agrowaste. BioResources. 2014a;9(2):3508-3525
  28. 28. Mohapatra D, Mishra S, Sutar N. Banana and its by-product utilisation: An overview. Journal of Scientific and Industrial Research. 2010;69:323-329
  29. 29. ASTM. Standard Test Method for Kinematic Viscosity of Transparent and Opaque Liquids (the Calculation of Dynamic Viscosity). West Conshohocken, PA: ASTM International; 1997. www.astm.org
  30. 30. Ebenso E, Eddy N, Odiongenyi A. Corrosion inhibition and adsorption properties of methocarbamol on mild steel in acidic medium. Portugaliae Electrochimica Acta. 2009;27(1):13-22
  31. 31. Okafor P, Ebenso E. Inhibitive action of Carica papaya extracts on the corrosion of mild steel in acidic media and their adsorption characteristics. Pigment & Resin Technology. 2007;36(3):134-140
  32. 32. Maneerat N, Tangsuphoom N, Nitithamyong A. Effect of extraction condition on properties of pectin from banana peels and its function as fat replacer in salad cream. Journal of Food Science and Technology. 2017;54(2):386-397
  33. 33. SPE. Oil and gas facilities: Corrosion monitoring and mitigation technologies. Society of Petroleum Engineering. 2013;6(2):1-6
  34. 34. Kliškić M, Radošević J, Gudić S, Katalinić V. Aqueous extract of Rosmarinus officinalis L. as inhibitor of Al–Mg alloy corrosion in chloride solution. Journal of Applied Electrochemistry. 2000;30(7):823-830
  35. 35. Betiku E, Akintunde AM, Ojumu TV. Banana peels as a biobase catalyst for fatty acid methyl esters production using Napoleon's plume (Bauhinia monandra) seed oil: A process parameters optimization study. Energy. 2016;103:797-806
  36. 36. Noor EA, Al-Moubaraki AH. Corrosion behavior of mild steel in hydrochloric acid solutions. International Journal of Electrochemical Science. 2008;3(1):806-818
  37. 37. Gunavathy N, Murugavel S. Corrosion inhibition studies of mild steel in acid medium using Musa acuminata fruit peel extract. Journal of Chemistry. 2012;9(1):487-495
  38. 38. Mayanglambam RS, Sharma V, Singh G. Musa Paradisiaca extract as a green inhibitor for corrosion of mild steel in 0.5 M sulphuric acid solution. Portugaliae Electrochimica Acta. 2011;2011, 29(6):405-417
  39. 39. Lai C, Xie B, Zou L, Zheng X, Ma X, Zhu S. Adsorption and corrosion inhibition of mild steel in hydrochloric acid solution by S-allyl-O, O′-dialkyldithiophosphates. Results in Physics. 2017;7:3434-3443
  40. 40. Amodu OS, Ntwampe SK, Ojumu TV. Optimization of biosurfactant production by Bacillus licheniformis STK 01 grown exclusively on Beta vulgaris waste using response surface methodology. BioResources. 2014b;9(3):5045-5065
  41. 41. Akalezi CO, Enenebaku CK, Oguzie EE. Inhibition of acid corrosion of mild steel by biomass extract from the Petersianthus macrocarpus plant. Journal of Materials and Environmental Science. 2013;4(2):217-226
  42. 42. Raja PB, Qureshi AK, Rahim AA, Osman H, Awang K. Neolamarckia cadamba alkaloids as eco-friendly corrosion inhibitors for mild steel in 1 M HCl media. Corrosion Science. 2013;69:292-301
  43. 43. Negm N, Kandile N, Aiad I, Mohammad M. New eco-friendly cationic surfactants: Synthesis, characterization and applicability as corrosion inhibitors for carbon steel in 1 N HCl. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2011;391(1-3):224-233
  44. 44. Gunasekaran G, Chauhan L. Eco friendly inhibitor for corrosion inhibition of mild steel in phosphoric acid medium. Electrochimica Acta. 2004;49(25):4387-4395
  45. 45. Verma C, Quraishi M, Ebenso EE, Bahadur I. A green and sustainable approach for mild steel acidic corrosion inhibition using leaves extract: Experimental and DFT studies. Journal of Bio-and Tribo-Corrosion. 2018;4(3):33
  46. 46. Ong CC, Karim KA. Inhibitory effect of red onion skin extract on the corrosion of mild steel in acidic medium. Chemical Engineering Transactions. 2017;56:913-918
  47. 47. Bao J, Zhang H, Zhao X, Deng J. Biomass polymeric microspheres containing aldehyde groups: Immobilizing and controlled-releasing amino acids as green metal corrosion inhibitor. Chemical Engineering Journal. 2018;341:146-156

Written By

Olusola S. Amodu, Moradeyo O. Odunlami, Joseph T. Akintola, Seteno K. Ntwampe and Seide M. Akoro

Submitted: 25 July 2018 Reviewed: 20 November 2018 Published: 12 February 2019