Some alginates polymeric cationic surfactants

0 downloads 0 Views 2MB Size Report
Oct 7, 2018 - converting the natural alginic acid into polymeric cationic surfactant, and their evaluating as ... Contents lists available at ScienceDirect. Journal of ... acid monomer. (17.6 g) with 0.1 mol from 2(2(dimethylamino)ethoxy)ethanol.
Journal of Molecular Liquids 273 (2019) 164–176

Contents lists available at ScienceDirect

Journal of Molecular Liquids journal homepage: www.elsevier.com/locate/molliq

Some alginates polymeric cationic surfactants; surface study and their evaluation as biocide and corrosion inhibitors Samy M. Shaban a,b,⁎, I. Aiad a, Ahmed H. Moustafa c, Omar H. Aljoboury c a b c

Petrochemical Department, Egyptian Petroleum Research Institute, Egypt School of Chemical Engineering, Sungkyunkwan University, 16419, Republic of Korea Chemistry Department, Faculty of Science, Zagazig University, Zagazig, Egypt

a r t i c l e

i n f o

Article history: Received 7 August 2018 Received in revised form 19 September 2018 Accepted 3 October 2018 Available online 07 October 2018 Keywords: Alginate cationic polymeric surfactants Surface parameters Steel corrosion Emulsion stability Biological activity

a b s t r a c t Three cationic polymeric surfactants were synthesized by alkylation of prepared alginate ester to obtain on three polymers with different alkyl chain (octyl, dodecyl and hexadecyl), and they are labeled (ALGOB, ALGDB and ALGHB), respectively. The chemical structure was confirmed using FTIR and 1HNMR spectroscopy. The behavior of the synthesized polymeric cationic surfactants in aqueous solution has been determined at three temperature 25, 45 and 65 °C depending on surface tension and conductivity measurements. The emulsion stability, interfacial tension and foaming power of the synthesized cationic surfactant were studied. The efficiency of the ALGOB, ALGDB and ALGHB inhibitor against the corrosion of mild steel in the aggressive solution (1.0 M HCl) were evaluated gravimetrically at various temperatures 25, 40, 55 and 70 °C. The corrosion inhibition efficiency (ղ) has been increased with the elongation of the hydrocarbon chain length and with raising the solution temperature. The maximum ղ achieved was equal to 92.6 for the synthesized polymeric surfactant ALGHB at temperature 70 °C. The adsorption Villamil isotherm is the best isotherm describing the adsorption of ALGOB, ALGDB and ALGHB polymeric surfactant on the mild steel surface. The electrochemical study of the behavior of the synthesized polymeric surfactants outlined that they behave as mixed type inhibitor (Tafel curves) and thickness of the adsorbed layer on the steel surface increase as the double layer capacitance decrease (Impedance study). The prepared polymeric surfactants showed good activity against some common G (+ve) and G (−ve) bacteria as well as Fungi. © 2018 Elsevier B.V. All rights reserved.

1. Introduction The surfactant compounds spread around the world and attracted the attention of the scientist due to their unique structure. The surfactant materials comprise of two opposing parts, one is hydrophilic head and the other is the tail, which is hydrophobic [1–3]. The surfactants are characterized by their ability to aggregate in cluster forms in solutions, which is a vital process in many applications like pharmaceutical applications, detergent, designing the microemulsion, and emulsion polymerization [4–6]. At the same time, the surfactants characterized by their tendency to adsorb on the interfaces, which is a very important process in different application like corrosion inhibitors, bactericide, emulsifier and enhancing oil recovery [7–10]. The surfactants used in many research fields during the preparation of nanomaterial, catalyst, sensors and drug deliver [11,12]. The carbon steel considers the main, constructing part in the pipelines and process involved in the petroleum ⁎ Corresponding author. E-mail addresses: [email protected], [email protected] (S.M. Shaban).

https://doi.org/10.1016/j.molliq.2018.10.017 0167-7322/© 2018 Elsevier B.V. All rights reserved.

oil and gas production sectors; consequently their protection represents a serious case [13,14]. The choice of the suitable corrosion inhibitors, mainly depends on their chemical structure to contain electronic rich functional group like oxygen, nitrogen, double bond, triple bond and aromatic acting as an active center during the adsorption process [15]. The surfactants corrosion inhibitors are used in large extend in protection the steel in oil petroleum sectors due to their high adsorption affinity at the interfaces. The surfactant hydrophobic tail form layers which consider as a barrier between the aggressive medium and the steel [16]. The alginates derivatives have been used in pharmaceutical and biomedical applications due to their low toxicity, low cost, biocompatibility and biodegradation where it extracted from brown seaweed [17]. Our work aimed to prepare polymeric cationic surfactants by converting the natural alginic acid into polymeric cationic surfactant, and their evaluating as corrosion inhibitors for the steel in 1.0 M HCl in an aggressive medium depending on gravimetric and electrochemical methods. Also, the surface parameters of the compounds have been assessed depending on surface tension and conductometric measurements. The activity of the materials as antimicrobial against the some

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

pathogenic bacteria and fungi has been assessed using disc diffusion method. 2. Materials and experimental techniques 2.1. Materials The used chemicals in the preparation of the polymeric cationic surfactants are 2 2 (2 (dimethylamino)ethoxy)ethanol, Octyl bromide, Dodecyl bromide, and Hexadecyl bromide (purchased from SigmaAldrich) and alginic acid (purchased from Arcos company). 2.2. Synthesis of cationic polymeric surfactant The reaction proceeds in two steps:

165

B. Alginate ester alkylation

The laboratory obtained alginate ester from the previous step (0.015 M, 4.64 g) was added to (0.015 mol) from the fatty alkyl bromide namely (hexadecyl bromide (4.58 g), dodecyl bromide (3.74 g) and octyl bromide (2.89 g)) in 250 mL round flask. To the previous reaction mixture, 100 mL from the absolute ethanol was added and refluxed for 25–30 h. The reaction was monitored using thin layer chromatography and FTIR, after that the solvent was evaporated in a dry oven and then the obtained residual has been crystallized using acetone and diethyl ether [19]. The obtained polymeric surfactant with the hydrophobic tail equal to eight, twelve and sixteen carbon atoms are labeled ALGOB, ALGDB and ALGHB with yield percent 93, 94.5 and 95.2% respectively. The reaction pathway has been represented in the Scheme 1.

A. Esterification of alginic acid 2.3. Structure characterization The ester is formed by reaction 0.1 mol the alginic acid monomer (17.6 g) with 0.1 mol from 2 (2 (dimethylamino)ethoxy)ethanol (13.3 g) in (150 mL) xylene. A 0.01% from the p toluene sulphonic acid catalyst (PTSA) was added to the reaction mixture. The reaction was finished when 1.8 mL of water was received in the Dean-Stark apparatus. The solvent was evaporated in a dry oven at 120 °C, and then petroleum ether was used to get rid of the PTSA from the obtained alginate ester with 90% yield [18].

The reaction pathway progress was monitored through Fourier transform infrared spectroscopy and 1H NMR spectroscopy. The ATI Mattsonm Infinity Series™ instrument (Bench top 961 incorporated in the Win First™ V2.01 Software) in the EPRI Institute, Egypt was used for detecting the functional group of the synthesized polymeric surfactant. The 1H NMR was performed in Faculty of Pharmacy - Cairo University using (Bruker High Performance Digital FT-NMR Spectrometer A Vance III 400 MHz).

Scheme 1. Synthetic pathway of polymeric cationic surfactants.

166

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

stopping the shaking. The stability the produced foam is represented by the consumed time for the foam height to be halved [27].

2.4. Measurements 2.4.1. Specific conductivity The critical micelle concentration (CMC) of the ALGOB, ALGDB and ALGHB polymeric surfactant at 25, 45 and 65 °C, were estimated from the specific conductivity measurements which provided using Cond 3210 SET 1, Probe tetra corn 325 (Wissenschaftlich Technische Werkstattern) apparatus. The distilled water has been used for preparing different concentration from ALGOB, ALGDB and ALGHB surfactant starting from 0.01 to 2000 ppm. Water bath was used to adjust the temperature [20,21]. The CMC corresponds to the inflection point in the specific conductivity-concentration plot. 2.4.2. Surface and interfacial tension measurements (γ) The surface tension at 25, 45 and 65 °C of the aqueous solution from the prepared cationic polymeric surfactant ALGOB, ALGDB and ALGHB were performed using tensiometer-K6 instrument (KrÜss Company, Germany), the used method is the ring method. The distilled water has been used for calibrating and for preparation different concentration from ALGOB, ALGDB and ALGHB. Different concentrations from 0.01 to 2000 ppm have been prepared and their surface tension has been determined. Each concentration was repeated triplicate and the average was then taken. The interfacial tension was performed between light paraffin oil and 0.1 g/L from ALGOB, ALGDB and ALGHB polymeric surfactant at 25 °C using the same instrument KrÜss K-6 [22,23]. The inflection point in the surface tension (γ) versus [log c] plots of the synthesized polymeric surfactant ALGOB, ALGDB and ALGHB represent the critical micelle concentrations (CMC) of the ALGOB, ALGDB and ALGHB correspond to [24,25]. The ALGOB, ALGDB and ALGHB surfactant concentration capable to minimize the blank surface tension by 20 mN/cm is known by the surfactant efficiency (C20). The effectiveness (πCMC) of the synthesized cationic polymeric surfactant ALGOB, ALGDB and ALGHB equal to the difference in the surface tension of blank water (γo) and that at the CMC (γCMC) as described in the following specified Eq. (1): πCMC ¼ γ o −γCMC

ð1Þ

The maximum surface excess (Γmax) is the concentration of the ALGOB, ALGDB and ALGHB polymeric surfactants unimers at the phase interface per unit area and has been calculated according to Gibb's adsorption Eq. (2) [26].  Γ max ¼

1 2:303 nRT



δγ δ log c

 ð2Þ T

The R, represent the universal gas constant, the symbol n, represents the of active species number (n equal 2 for cationic amphipathic), and T is the absolute solution temperature. The average area in square angstrom occupied by the adsorbed ALGOB, ALGDB and ALGHB polymeric surfactant unimers at the phase interface is known by the minimum surface area (Amin). The Amin has been measured through the Gibb's adsorption Eq. (3), where the symbol N is equal to Avogadro's number [26].

Amin

1016 ¼ Γ max N

2.4.4. Emulsion stability The emulsion stability of the ALGOB, ALGDB and ALGHB surfactant was measured by strongly shacking a solution mixture comprise from 10 mL of the polymeric surfactant aqueous solution (0.1%) and 10 mL of the light paraffin oil at 25 °C in 50 mL measuring cylinder with a stopper. The time required for separation of (9 mL) of polymeric surfactant solution is the actually experimental emulsion stability [28]. 2.4.5. Corrosion measurement The affinity of the synthesized polymeric corrosion inhibitors ALGOB, ALGDB and ALGHB to inhibit the steel corrosion in 1.0 M HCl has been determined electrically at 25 °C and gravimetrically at 25, 40, 55 and 70 °C. The weight loss experiments were performed using mild steel specimens with fixed dimension of 3 cm × 6 cm × 0.4 cm. Each specimen has been abraded using emery papers with different grade from 400 to 1200, then it degreased with acetone and dried before weighting using a digital balance. The specimens have been immersed in the corrosive solution either with or without the synthesized ALGOB, ALGDB and ALGHB polymeric inhibitor for 24 h as immersion time. After 24 h, the specimens were removed and washed, then dried before reweighting again. A water bath with thermostat was used during the experiment to control the temperature [29,30]. Regarding to the electrochemical measurements, they were performed using a Voltalab 40 Potentiostat instrument of PGZ 301 model combined with Voltamaster 4 software. Glass cell with three necks has been used and connected with a platinum counter electrode and a saturated calomel electrode (SCE). The exposed area of carbon steel working electrode is 0.7 cm2. The exposed surface was treated as in weight loss experiment. The Tafel curves were obtained by changing the applied potential automatically from −800 to −200 mV with a scan rate equal to 0.1 mV s−1 [31]. The electrochemical impedance measurements were performed in a frequency range of 100 kHz to 50 MHz with a 5 mV [32]. The chemical composition of the working electrode, which used in weight loss and electrochemical measurement are the same, which is C (0.17%), P (0.02%), Mn (0.472%), Si (0.25%). S (0.018%) and the rest percent is iron. 2.4.6. Biological activity The biological activity of the three prepared polymeric surfactants ALGOB, ALGDB and ALGHB was assessed using a modified Kirby-Bauer disc diffusion method against some of the Gram positive G (+ve), Gram negative G (−ve) pathogenic bacteria and fungi [33]. The pathogenic G (+ve) bacteria are Bacillus subtilis (ATCC 6051), and Staphylococcus aureus (ATCC 12600), while the pathogenic G (−ve) are Escherichia coli (ATCC 11775), and Pseudomonas aeruginosa (ATCC 10145). The tested fungi are Candida albicans (ATCC 7102) and Aspergillus flavus. The test was performed in the micro analytical center (Cairo University, Egypt). The evaluation procedures were outlined in details in previous work [34]. 3. Results and discussion 3.1. Structure elucidation of the synthesized polymeric cationic surfactant

ð3Þ

2.4.3. Foam power The Foam power of the ALGOB, ALGDB and ALGHB polymeric surfactant were determined by strongly shaking of 100 mL of their aqueous solution (0.1%) in a graduated cylinder with a stopper (250 mL) for 5 min at 25 °C. The foam power of the ALGOB, ALGDB and ALGHB polymeric surfactant is the initially produced foam height (in mL) after

The reaction pathway of the synthesized polymeric cationic surfactant ALGOB, ALGDB and ALGHB was confirmed using the FTIR spectroscopy. Fig. 1 describes the FTIR of the synthesized polymeric cationic surfactant ALGOB as an example, which refer to appearing a new carbonyl ester group at 1739 cm−1 and dispatching the carboxylic carbonyl group. Some characteristic bands have been initiated like that at 2928 and 2852 cm−1 for asymmetric and symmetric stretching aliphatic (CH), respectively. In addition to, the bands at 1472 and 1400 cm−1 characterized for the bending aliphatic CH2 and CH3, respectively.

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

167

Fig. 1. FTIR spectrum of the polymeric cationic surfactant ALGOB. 1

H NMR spectrum of the synthesized ALGDB (Fig. 2), revealed signals at: δ = 0.86 (t, 3H, CH3, alkyl chain), 1.07 (m, 2H, CH2(CH2)8 CH2CH3), 1.24 (m,16H, \\CH2(CH2)8CH2CH3), 1.66 (m, 2H, \\CH2CH2 (CH2)8CH2CH3), 2.51 (t, 2H,\\CH2 CH2(CH2)8CH2CH3), 2.8–2.81 (s, 6H, \\CH2N⊕(CH3)2CH2\\), 3.06 (t, 2H, \\OCH2CH2N⊕(CH3)2CH2\\), 3.27–3.41 (m, 5H, alginic ring protons), 3.5–3.51 (d, 2H, 2OH for alginic ring), 3.53 (t, 3H, \\OCH2CH2N⊕(CH3)2CH2\\), 3.54 (t, 3H, \\OCH2CH2OCH2CH2N⊕(CH3)2CH2\\) and at 3.73 (t, 3H, \\OCH2CH2OCH2CH2N⊕(CH3)2CH2\\). 3.2. Surface study The effect of increasing the hydrophobicity of the prepared polymeric surfactants ALGOB, ALGDB and ALGHB on the solution conductivity at 45 °C was outlined in Fig. 3, while, the effect of raising the solution temperature from 25 to 65 °C was outlined in Fig. 4. The experimental results referred to decreasing the solution specific conductivity with

increasing the hydrophobic chain length of the synthesized polymeric cationic surfactant ALGOB, ALGDB and ALGHB, as it's obvious in (Fig. 3). The specific conductivity increased also by raising the solution temperature as it is clear in the (Fig. 4), for the synthesized ALGOB as an example. The calculated degree of counter ion dissociation values (α) for the synthesized polymeric cationic surfactant ALGOB, ALGDB and ALGHB at 25, 45 65 °C were depicted in Table 1. The calculated (α) values give insight on the presence of a reversible relationship between the conductivity and the hydrophobic chain length. The (α) values of ALGOB, ALGDB and ALGHB at 60 °C are equal to 0.63, 0.59 and 0.58 respectively. Furthermore, the calculated α value refers to presence a direct relationship with elevating the solution temperature up to 65 °C. The α values of the ALGOB at 25, 45 and 65 °C are equal to 0.55, 0.60 and 0.63, respectively. With elongation the hydrophobic tail length, the amount of the hydrated water around the hydrophilic head of the ALGOB, ALGDB and

Fig. 2. 1HNMR spectrum of the polymeric cationic surfactant ALGDB.

168

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

Fig. 3. The plots of specific conductivity against concentrations of the prepared polymeric cationic surfactants in distalled water at 45 °C.

ALGHB unimers or their micelles are decreased, which enhance the bond strength between the heads and their associated counter ions, leading to increasing the experimental α values (Table 1) [35,36]. Raising the solution temperature lead to weakness the bond strength between the ALGOB, ALGDB and ALGHB head and its associated counter ion [37], consequently increasing the α values as it is outlined in Table 1 and Fig. 4. The surface parameters of the synthesized ALGOB, ALGDB and ALGHB including critical micelle concentration, efficiency, effectiveness, minimum surface area and maximum surface excess were determined based on surface tension and conductivity measurements at 25, 45 and 65 °C and the obtained results were outlined in Table 1. The break point in the surface tension- log[C] curves for the synthesized ALGOB, ALGDB and ALGHB surfactants represent the CMC, Figs. 5–7. The CMCs values of the ALGOB, ALGDB and ALGHB were determined also from the specific conductivity data and represent the inflection in the specific conductivity versus aqueous surfactant concentration curves (Figs. 3 & 4). There is an instance increase in the specific conductivity values at the point where the micelles start to be formed as it is outlined in Figs. 3 & 4, which ascribed to increasing the water hydration around the micelles head, which decrease the bond strength with the associated counter ions, hence increasing the specific conductivity. The experimental CMCs values in Table 1 and Fig. 8, refer to their decreasing with elongation the hydrophobic tail ALGOB, ALGDB and ALGHB surfactant. This trend is ascribed to destroying the structure of

Fig. 4. The plots of specific conductivity against concentrations of the prepared polymeric cationic surfactant (ALGOB) in distalled water at 25, 45 and 65 °C.

the water with increasing the hydrophobicity, consequently the free energy of the aqueous surfactant solution will increase and allowing the unimers to aggregate together in micelles forms in a way to minimize the up increase in the free energy, Consequently, increasing their affinity to aggregates in micelles at lower concentration [38,39]. For example, the critical micelle concentration values of the synthesized polymeric cationic surfactant ALGOB, ALGDB and ALGHB at 65 °C, are equal to 740.6, 637.3 and 280.9 ppm, respectively (surface tension measurements, Table 1). The solution temperature has a marked effect on the characteristic CMCs values of ALGOB, ALGDB and ALGHB polymeric surfactants at 25, 45 and 65 °C. The experimental CMC values of the ALGOB, ALGDB and ALGHB polymeric surfactants decrease with increasing the solution temperature (Fig. 8). The CMCs values ALGHB polymeric surfactants are equal to 789, 419.16 and 298.5 ppm at 25, 45 and 65 °C, respectively [surface tension measurements). Increasing the solution temperature, the amount of the hydrated water around the head and tail is decreased. Decreasing the water hydration around the hydrophilic head, enhance the micellization affinity, while, the decreasing around the tail decrease their affinity to form micelle [40,41]. The net magnitude of the two factors is responsible for the temperature effect on the CMC of the ALGOB, ALGDB and ALGHB surfactant. The experimental CMCs values in Table 1, outlined that the CMC values of the synthesized ALGOB, ALGDB and ALGHB decrease with raising the solution temperature. The efficiency (C20) values of ALGOB, ALGDB and ALGHB polymeric surfactant in Table 1, declared that the tail has a positive effect on the decreasing the surface tension values (increasing the efficiency). With increasing the surfactant tail, their migration to the interface will increase, which enhance their efficiency to decrease the surface tension [42]. For instance, the efficiency of ALGOB, ALGDB and ALGHB is equal to 119.8, 65.4 and 28.6 ppm at 25 °C respectively (Table 1). Increasing the ALGOB, ALGDB and ALGHB solution temperature, their affinity to decrease the surface tension is enhanced, as it clears from the depicted values in Table 1. The C20 values of the ALGOB polymeric surfactant are equal to 119.8, 105.13 and 75.9 ppm at 25, 45 and 65 °C, respectively. As outlined previously, the major effect of raising solution temperature is decreasing the water hydration around the ALGOB, ALGDB and ALGHB hydrophilic head, which enhance the surfactant migration to the solution interface consequently, more reduction in the surface tension [43]. The depicted effectiveness values πCMC in Table 1, refer to that the prepared ALGHB with longer tail is the most effective one. The πCMC values of the ALGOB, ALGDB and ALGHB surfactant at 45 °C equal to 31.9, 34.9 and 35.9 mN m−1, respectively (Table 1), where it has the higher CMC/C20 values at all the tested temperatures as it is outlined in Table 1. The CMC/C20 values of the prepared ALGOB, ALGDB and ALGHB surfactants were 7.9, 13.3 and 26.5 at temperature 25 °C, respectively [44,45]. The calculated Γmax and the Amin values of the synthesized ALGOB, ALGDB and ALGHB were depicted in Table 1. The experimental data revealed that increasing the hydrocarbon tail of the synthesized ALGOB, ALGDB and ALGHB is followed by increasing the amount of accumulated surfactant unimers at the phase interface, and consequently decreasing the occupied area by theses unimers at the interface. The Amin of the surfactants ALGOB, ALGDB and ALGHB is equal to 292.4, 271.1 and 259.04 A2, respectively at temperature 45 °C, while their Γmax is equal to 0.57, 0.61 and 0.64 (×10−10 mol cm−2), respectively. The longer tail of the surfactant ALGHB, enhance its adsorption tendency at the interface, consequently the amount of the accumulated surfactant unimer at the interface increase (Γmax increase). Hence, the formed adsorbed surfactant unimer monolayer expected to be highly dense [46,47]. The dense and compacted layer is important in several interfacial applications like emulsification, biological activity, corrosion inhibition and solubilization. The high accumulated adsorbed layer force the surfactant unimer to be more less perpendicular to the interface, hence the minimum area is decreased.

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

169

Table 1 The surface and thermodynamic properties of synthesized polymeric cationic surfactant at different temperatures. Comp.

Temp. oC

CMCa/ (ppm)

CMCb/ (ppm)

α

C20 (ppm)

πCMC/(mN m−1)

Гmax *10−10 (ppm·cm−2)

Amin/A2

CMC/C20

ALGOB

25 45 65 25 45 65 25 45 65

954.99 904.74 740.57 871.40 740.57 637.31 759.34 406.16 280.93

975.76 912.33 748.72 893.15 764.44 657.58 789.01 419.57 298.55

0.55 0.60 0.63 0.53 0.57 0.59 0.51 0.54 0.58

119.85 105.13 75.93 65.35 41.01 29.34 28.61 9.05 7.40

31.99 31.93 28.78 35.31 34.91 32.84 37.36 35.99 33.75

0.64 0.57 0.48 0.70 0.61 0.52 0.72 0.64 0.55

259.83 292.41 347.74 236.84 271.12 317.37 230.08 259.04 300.98

7.97 8.61 9.75 13.34 18.06 21.72 26.54 44.89 37.97

ALGDB

ALGHB

a b

Surface tension measurements. Conductivity measurements.

where W is the values of the weight loss in the presence of the synthesized polymeric cationic inhibitor, while Wο is the weight loss in their

absence. The ΔW is the average weight loss, S is the total surface area of the coupon in cm2 and t refer to the contact immersion time in hours. The experimental inhibition efficiency (ղw) and corrosion rate (k) for the ALGOB, ALGDB and ALGHB at 25, 40, 55 and 70 °C were depicted in Table 2. The ղw increasing with raising the inhibitor concentration, is due to increasing the amount of the accumulated ALGOB, ALGDB and ALGHB surfactant at the interface, hence the covered area on the steel surface has been increased. Which act as an insulator between the aggressive ions and the steel surface [48]. The experimental dataset in Table 2 and Fig. 9, point to increasing the inhibition efficiency with increasing the hydrophobic chain length of the ALGOB, ALGDB and ALGHB, where their ղw at 70 °C, are equal to 81.8, 85.8 and 89.28% at concentration 500 ppm, respectively. As discussed in the previous section, increasing the surfactant chain length, the adsorption affinity at the interface increase, at the same time the maximum surface excess was achieved with the surfactant ALGHB (Table 1), hence covered area increased and more inhibition efficiency is obtained. The inhibiting effect of the steel corrosion in the presence of ALGOB, ALGDB and ALGHB was increased with raising the solution temperature from 25 to 70 °C, as it is depicted in Table 2 and Fig. 9. The ղw of the ALGDB at 25, 40, 55 and 70 °C were equal to 71.8, 79.6, 83.3 and 85.8% respectively, at 500 ppm. Raising the solution temperature enhance the electrochemical reactions through increasing the activation energy of the reacted species. This behavior refers to presence a chemisorption between the ALGOB, ALGDB and ALGHB through their forming coordinate bonds between the vacant d-orbital's of the iron and high rich electronic cloud of the inhibitors (unshared lone pairs of electrons on the N and O atoms) [49]. Raising the temperature of the inhibited solution, some chemical changes were occurred leading to strong adsorption to the metal surface through the active groups in the synthesized alginic polymeric inhibitors [50].

Fig. 5. Surface tension and Log concentration relationship of synthesized polymeric cationic surfactant (ALGOB) at different temperatures.

Fig. 6. Surface tension and Log concentration relationship of synthesized polymeric cationic surfactant (ALGDB) at different temperatures.

Increasing the solution temperature, increases the activation energy as well as decreasing the hydration around both hydrophilic head and hydrophobic tail, which enhance the ALGOB, ALGDB and ALGHB migration to the interface. By decreasing the hydration around the hydrophilic head, the repulsion force between positively charged surfactant heads increase, which force the unimers to be less perpendicular to the interface, consequently the area occupied is increased. The increase in the minimum surface area, with temperature, lead to decreasing the packing density of the accumulated surfactant unimer at the interface hence decreasing the maximum surface excess. 3.3. Corrosion investigation I. Gravimetric study

The experimental corrosion rate (k), surface coverage (θ) and the inhibition efficiency (ɳw) have been obtained according to Eqs. (4), (5) and (6) depending on the weight loss measurements: k¼

ΔW St

ð4Þ

θ¼

W ο −W Wο

ð5Þ

ηw ¼

W ο −W  100 Wο

ð6Þ

170

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

Plotting ln value (K / T) versus value 1 / T, we obtained on straight lines as described in the Fig. S2 (supplementary). The slope are equal to ΔH* / R and the intercept refer to log ((R / NAh) + ΔS* / R). The depicted ΔH* and ΔS* values in Table 3, reflect the endothermic nature for the steel dissolution process in 1.0 M HCl as aggressive medium. The ΔH* values are positive, which indicates the difficulty of the steel dissolution in the presence of the synthesized ALGOB, ALGDB and ALGHB [53]. The negative values of the entropy change of activation (Table 3), refer that the activated complex in the rate-determining step is an association rather than dissociation [54]. The depicted weight loss data in Table 2, have been used for fitting different adsorption isotherm, for detecting the most suitable one describing the adsorption mechanism of the synthesized ALGOB, ALGDB and ALGHB inhibitors on the steel surface in 1.0 M HCl at 25, 40, 55 and 70 °C. The best-fitted isotherm is Langmuir isotherm as described in Eq. (13) and outlined in Fig. S3 (supplementary) with linear regression coefficient close to 1. 



Fig. 7. Surface tension and Log concentration relationship of synthesized polymeric cationic surfactant (ALGHB) at different temperatures.

C ¼ θ

The kinetics of the corrosion process have been assessed using the following Arrhenius equation

where, Kads is the equilibrium constant and C is the ALGOB, ALGDB and ALGHB concentration in ppm. The obtained slopes (Table 4), were found more than 1, referring that each ALGOB, ALGDB and ALGHB unimer occupy more than one adsorption center, which represent deviation from the Langmuir postulates. Consequently, the Langmuir has been modified to Villamil isotherm (modified Langmuir isotherm), as described in Eq. (14) [55].

ln K ¼

−Ea þ lnA RT

ð7Þ

where, Ea is the activation energy, while K and A are the corrosion rate and the pre- Arrhenius constant, respectively. Fig. S1 (supplementary) describe the Arrhenius plots of Ln K vs. 1 / T for the prepared ALGOB, ALGDB and ALGHB respectively. All the plots give a straight line with linear regression coefficient nearly equal 1, as a confirmation for obeying the Arrhenius equation. The Ea was obtained from slope, which a equal to ( −E RT ). The Ea of the uninhibited solution are equal to 43.14 kJ/mol, while, for the inhibited solution with 500 ppm from ALGOB, ALGDB and ALGHB inhibitor were equal to 32.8, 30.2 and 27.5 kJ/mol, respectively. The lower Ea values of the inhibited solution could be attributed to the presence of chemisorption of prepared polymeric surfactants on the mild steel surface, which is supported by their bigger size [51,52]. The activation enthalpy change and entropy (ΔH*, ΔS*) values were calculated according to the transition state theory Eq. (8): ln

    K R ΔS ΔH  ¼ ln þ − T NA h R RT

C ¼ θ

1 K ads



n K ads

þC

ð13Þ

 þ nC

ð14Þ

where n refers to the number of displacements water molecule adsorbed on the steel. The standard free energy of ALGOB, ALGDB and ALGHB surfactant adsorption (ΔGoads) has been calculated depending on the Kads values which obtained according to the Eq. (15) [56]. ° ¼ −RT ln ðαK ads Þ ΔGads

ð15Þ

The parameter α, represents the water concentration in solution (1 × 106 ppm), the constant R, is the universal gas constant while the T, is the absolute temperature. The adsorption heat (ΔHoads) values have been calculated using Van't Hoff Eq. (16).

ð8Þ ln K ads ¼

° −ΔHads RT

! þ Constant

ð16Þ

The standard adsorption entropy (ΔSoads) values have been calculated using the Eq. (17). ° ° ° ΔGads ¼ ΔH ads −TΔSads

Fig. 8. Hydrophobic chain length effect on the critical micelle concentration of synthesized polymeric cationic surfactants.

ð17Þ

The all calculated ΔGoads, ΔHoads and ΔSoads for the synthesized ALGOB, ALGDB and ALGHB were depicted in Table 4. The negative values of ΔGoads from −26.59 to −32.09 kJ/mol−1, suggesting that the spontaneous adsorption process of the ALGOB, ALGDB and ALGHB on the metal surface in 1.0 M HCl at 25, 40, 55 and 70 °C, are mixed between chemical and physical adsorption. The positive ΔHoads values of ALGOB, ALGDB and ALGHB polymeric inhibitors, which equal to 3.96, 7.43 and 9.12 kJ/mol, respectively, refer to the endothermic process of the synthesized ALGOB, ALGDB and ALGHB on the steel surface. The positive sign of ΔS°ads is attributed to the solvent entropy increasing and more water desorption entropy [57]. The electronic rich adsorption center in the synthesized polymeric surfactant ALGOB, ALGDB and ALGHB are oxygen, nitrogen, hydroxyl,

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

171

Table 2 Corrosion rate of carbon steel, surface coverage and percentage inhibition efficiency for carbon steel in different concentration of prepared alginic surfactants at different temperatures. Temp. °C

inhibitor Conc, ppm

ALGOB Weight loss (mg)

K, mg cm−2 h−1

Ө

η,%

Weight loss (mg)

K, mg cm−2 h−1

Ө

η,%

Weight loss (mg)

K, mg cm−2 h−1

Ө

η,%

25

0 10 25 50 100 250 500 1000 0 10 25 50 100 250 500 1000 0 10 25 50 100 250 500 1000 0 10 25 50 100 250 500 1000

400 216 196.5 183.7 166.2 144.3 126.1 104.9 1023.2 487.2 447.5 403.5 356.2 301.2 244.3 192.2 2287 1020.4 887.6 820.4 719.4 592.7 466.1 358.6 3825 1629.8 1405 1234.7 1060.9 892.6 695.6 526.9

0.3561 0.1923 0.1749 0.1636 0.1480 0.1285 0.1123 0.0934 0.9110 0.4338 0.3984 0.3592 0.3171 0.2682 0.2175 0.1711 2.0361 0.9085 0.7902 0.7304 0.6405 0.5277 0.4150 0.3193 3.4054 1.4510 1.2509 1.0993 0.9445 0.7947 0.6193 0.4691

– 0.4600 0.5087 0.5407 0.5845 0.6392 0.6847 0.7377 – 0.5238 0.5626 0.6056 0.6519 0.7056 0.7612 0.8122 – 0.5538 0.6119 0.6413 0.6854 0.7408 0.7962 0.8432 – 0.5739 0.6327 0.6772 0.7226 0.7666 0.8181 0.8622

– 46.00 50.87 54.07 58.45 63.92 68.47 73.77 – 52.38 56.26 60.56 65.19 70.56 76.12 81.22 – 55.38 61.19 64.13 68.54 74.08 79.62 84.32 – 57.39 63.27 67.72 72.26 76.66 81.81 86.22

400 200.8 185.6 174.9 155.9 132.5 112.9 92.4 1023.2 471.1 412.8 370.1 325.2 263.3 208.9 165.5 2287 915.8 824.4 750.5 632.3 494.1 381.6 297.3 3825 1449.9 1256.3 1115.3 934.1 724.3 543 409

0.3561 0.1788 0.1652 0.1557 0.1388 0.1180 0.1005 0.0823 0.9110 0.4194 0.3675 0.3295 0.2895 0.2344 0.1860 0.1473 2.0361 0.8153 0.7340 0.6682 0.5629 0.4399 0.3397 0.2647 3.4054 1.2909 1.1185 0.9930 0.8316 0.6449 0.4834 0.3641

– 0.4980 0.5360 0.5627 0.6102 0.6687 0.7177 0.7690 – 0.5396 0.5966 0.6383 0.6822 0.7427 0.7958 0.8383 – 0.5996 0.6395 0.6718 0.7235 0.7840 0.8331 0.8700 – 0.6209 0.6716 0.7084 0.7558 0.8106 0.8580 0.8931

– 49.80 53.60 56.27 61.02 66.87 71.77 76.90 – 53.96 59.66 63.83 68.22 74.27 79.58 83.83 – 59.96 63.95 67.18 72.35 78.40 83.31 87.00 – 62.09 67.16 70.84 75.58 81.06 85.80 89.31

400 188.3 170.3 156.3 139 117.9 97.3 73.6 1023.2 431.8 378.1 337.4 297.3 233.7 181.4 141.3 2287 852.6 731.1 639.1 539 403.8 302.9 217.3 3825 1298.1 1082.6 925.1 765.3 572 410 281.5

0.3561 0.1676 0.1516 0.1392 0.1238 0.1050 0.0866 0.0655 0.9110 0.3844 0.3366 0.3004 0.2647 0.2081 0.1615 0.1258 2.0361 0.7591 0.6509 0.5690 0.4799 0.3595 0.2697 0.1935 3.4054 1.1557 0.9639 0.8236 0.6814 0.5093 0.3650 0.2506

– 0.5293 0.5742 0.6093 0.6525 0.7053 0.7568 0.8160 – 0.5780 0.6305 0.6703 0.7094 0.7716 0.8227 0.8619 – 0.6272 0.6803 0.7206 0.7643 0.8234 0.8676 0.9050 – 0.6606 0.7170 0.7581 0.7999 0.8505 0.8928 0.9264

– 52.93 57.42 60.93 65.25 70.53 75.68 81.60 – 57.80 63.05 67.03 70.94 77.16 82.27 86.19 – 62.72 68.03 72.06 76.43 82.34 86.76 90.50 – 66.06 71.70 75.81 79.99 85.05 89.28 92.64

40

55

70

ALGDB

ALGHB

Fig. 9. Variation of inhibition efficiency and corrosion rate versus logarithm C of the prepared polymeric cationic surfactants at A: (25 °C), B: (40 °C), C: (55 °C) and D: (70 °C).

172

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

Table 3 Activation parameters values of carbon steel in 1.0 M HCl of different concentrations of the synthesized polymeric inhibitors. Inhibitor name

Conc. of inhibitor (M)

Ea (kJ mol−1)

Linear regression coefficient

ΔH* (kJ mol−1)

ΔS* (J mol−1 K−1)

ALGOB

0.00 10 25 50 100 250 500 1000 10 25 50 100 250 500 1000 10 25 50 100 250 500 1000

43.14 38.70 37.49 36.69 35.67 34.96 32.81 30.89 37.56 36.61 35.67 34.37 32.58 30.23 28.73 36.87 35.44 34.04 32.57 30.11 27.51 25.44

0.9927 0.994 0.993 0.9913 0.9901 0.993 0.9949 0.9975 0.9918 0.9914 0.9911 0.9915 0.993 0.9933 0.9915 0.99 0.9889 0.9881 0.9861 0.99 0.9885 0.9767

40.48 36.05 34.84 33.93 33.01 32.31 30.15 28.41 34.91 33.95 33.01 31.71 29.93 27.58 26.07 34.21 32.72 31.38 29.91 27.45 24.85 22.78

−117.22 −137.36 −142.12 −145.74 −149.66 −153.27 −161.75 −169.19 −141.63 −145.56 −149.30 −154.60 −161.99 −171.29 −178.00 −144.48 −150.34 −155.54 −161.39 −171.13 −181.47 −190.52

ALGDB

ALGHB

carbonyl groups and double bonds, which can interact with vacant dorbital of the steel immersed in the 1.0 M HCl. The presence of the hydrophobic carbon chain length, forms successive layers isolating the steel from contacting with the aggressive medium. II. Potentiodynamic polarization

The Potential/current curves for the tested mild steel behavior in the aggressive 1.0 M HCl with and without different concentration from the synthesized ALGOB, ALGDB and ALGHB inhibitors at 25 °C were graphically outlined in Fig. 10. The depicted electrochemical corrosion parameters in Table 5 were obtained from the extrapolation of the experimental. The degree of covered surface area (θ) by the synthesized ALGOB, ALGDB and ALGHB inhibitors and the percentage inhibition efficiency (ηp %) were calculated using the Eqs. (9) and (10), respectively. θ ¼ 1−

i iο

 ղp ¼

1−

ð9Þ i iο

  100

ð10Þ

Table 4 Thermodynamic parameters from Villamil adsorption isotherm on carbon steel surface in 1.0 M HCl in presence of ALGOB, ALGDB and ALGHB polymeric surfactants. Inhibitor name

Temp. °C

Slope

R2

Kads × 102 M−1

ΔGads kJ mol−1

ΔHads kJ mol−1

ΔSads J mol−1 K−1

ALGOB

25 40 55 70 25 40 55 70 25 40 55 70

1.35 1.22 1.18 1.15 1.29 1.18 1.14 1.11 1.22 1.15 1.10 1.07

0.9982 0.9985 0.9987 0.9991 0.9984 0.9991 0.9993 0.9994 0.9981 0.9992 0.9994 0.9995

4.56 4.82 5.28 5.93 4.68 5.39 6.14 6.74 4.71 5.89 6.75 7.68

−26.59 −28.08 −29.67 −31.36 −26.66 −28.36 −30.08 −31.72 −26.67 −28.60 −30.34 −32.09

3.96

102.48 102.32 102.49 102.93 114.02 114.00 114.02 113.82 120.06 120.46 120.25 120.11

ALGDB

ALGHB

7.34

9.12

Fig. 10. Polarization curves for mild steel in 1.0 M HCl in the absence and presence of various concentrations of (A): ALGOB, (B): ALGDB, and (C): ALGHB.

where, io and i are the corrosion current densities of unprotected and protected steel with ALGOB, ALGDB and ALGHB polymeric cationic surfactant, respectively. The experimental results (Table 5), refer to increasing (ղp) as the concentration of the ALGOB, ALGDB and ALGHB inhibitors increase. For example the inhibition efficiency of the ALGHB at 10, 50, 100, and 500 ppm are equal to 52.38, 63.22, 65.8 and 75.7% respectively. This could be ascribed to increasing the surface covered by the surfactant as the concentration increase. The obtained i values (Table 5), was inversely proportional to the inhibitor concentration and with increasing the length of the hydrophobic tail. Consequently, the ηp values are increased with increasing the concentration and the hydrophobicity. The values of ηp for the synthesized ALGOB, ALGDB and ALGHB at concentration 250 ppm are equal to 63.7, 65.9 and 70.2% respectively. As the hydrophobicity, the adsorption affinity of the prepared polymeric surfactant at the interface increase as it is outlined in the previous section, hence the surface coverage increase leading to enhancing the inhibition efficiency. We observed that the maximum shift in Ecorr/SCE doesn't exceed 85 mV, referring that the synthesized ALGOB, ALGDB and ALGHB behave as a mixed type inhibitor. The potentiodynamic polarization curves are quite similar; where both the anodic and the cathodic Tafel slope (βa) and (βc) are slight shift. This suggests that the electrochemical mechanism has not been changed even in the presence of the synthesized ALGOB, ALGDB and ALGHB inhibitors [58–61].

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

173

Table 5 Potentiodynamic polarization measurements of mild steel in 0.5 M HCl in the absence and presence of various concentrations of nonionic surfactants HTOPD, HTOPT and HTOPH at 298 K. Inhibitor name

Conc. of inhibitor (M)

-Ecorr mV (SCE)

Icorr mA cm−2

-βc mV dec−1

βa mV dec−1

Ө

η%

ALGOB

0.00 10 25 50 100 250 500 1000 10 25 50 100 250 500 1000 10 25 50 100 250 500 1000

−530 −505.5 −527.3 −521.6 −522 −505 −523.8 −515.7 −532.9 −539.6 −504.8 −537.6 −546 −531.3 −530.9 −498.3 −517.6 −499 −515.4 −507.1 −525.9 −536.5

0.3238 0.1726 0.1583 0.1466 0.1345 0.1176 0.1025 0.0842 0.162 0.1497 0.1409 0.1223 0.1102 0.0862 0.0745 0.1542 0.1407 0.1191 0.1106 0.0965 0.0786 0.0646

−185.5 −153.7 −145.2 −142.5 −140.7 −137.3 −118.1 −121.9 −153.3 −154.3 −170.2 −136.8 −144.4 −130.6 −135.6 −175 −142.7 −143 −144 −161.8 −142.8 −134.4

189 128.2 269.7 234.1 309.3 105.6 239.5 168.3 273.6 282.2 190.4 219.8 253.7 255.4 181.4 236.3 262.6 225 246.9 246.2 199.3 216.1

00 0.4670 0.5111 0.5473 0.5846 0.6368 0.6834 0.7400 0.4997 0.5377 0.5649 0.6223 0.6597 0.7338 0.7699 0.5238 0.5655 0.6322 0.6584 0.7020 0.7573 0.8005

00 46.70 51.11 54.73 58.46 63.68 68.34 74.00 49.97 53.77 56.49 62.23 65.97 73.38 76.99 52.38 56.55 63.22 65.84 70.20 75.73 80.05

ALGDB

ALGHB

III. Electrochemical impedance study (EIS)

Fig. 11 outlines the impedance responses of tested carbon steel in the three synthesized polymeric cationic surfactants ALGOB, ALGDB and ALGHB solutions at 25 °C. Some observations have been deduced from the obtained Nyquist plots; which are: (i) The nearly similar shapes of Nyquist plots before and after adding the ALGOB, ALGDB and ALGHB polymeric inhibitors, which refer that the synthesized inhibitors do not alter the corrosion reaction mechanism [61,62]. (ii) The obtained Nyquist curves are single and depressed semicircles (not perfect semicircles), this deviation is related to mild steel inhomogeneity, roughness, and distribution of active sites [63]. (iii) The diameter of the Nyquist plots is increased with the incremental addition of the tested polymeric inhibitors ALGOB, ALGDB and ALGHB. The simple Randle equivalent circuit model has been used for fitting the Nyquist diagrams. The equivalent circuit consists of the solution resistance Rs and capacitanceCdl. The Cdl is parallel to the charge transfer resistance Rct [64]. The obtained electrochemical kinetics parameters had been outlined in the Table 6. The capacitance of the produced double layer (Cdl) is calculated using Eq. (11).   f −Z″img ¼ 1=ð2πC dl Rct Þ

ð11Þ

The parameter Z″img s the imaginary resistance components of the impedance are maximized. The inhibition efficiency percentages (ηz%) have been calculated through Eq. (12). η¼



  Rct −Ro ct =Rct  100

Fig. 11. Nyquist diagrams for mild steel in 1.0 M HCl in the absence and presence of various concentrations from the synthesized polymeric cationic surfactants.

by the ALGOB, ALGDB and ALGHB polymeric inhibitors. This could be ascribed to the adsorption of the synthesized polymeric surfactant on the mild steel interface and replacing the water molecules from the active cite. As known that, increasing the hydrophobic chain length enhance adsorption affinity, hence a lower in the capacitance values is predicted as it is depicted in Table 6. The formed layer from the synthesized polymeric inhibitors resist the charge transfer and consequently lead to increasing the Rct as it clear from increasing the diameter of the obtained semicircle (Fig. 11). 3.4. Interfacial tension and emulsification power

ð12Þ

Rct, is the charge transfer resistance in the presence of ALGOB, ALGDB and ALGHB while R°ct is in their absence. On inspecting the obtained data in Table 6) we note increasing the charge transfer resistance, Rct values and decreasing their corresponding capacitance Cdl, as increasing the concentration and with elongation the hydrophobic chain length. The decreasing in the Cdl values with increasing the inhibitor concentration may be due to decreasing the local dielectric constant and/or increasing the thickness of the electrical double layer formed

The synthesized polymeric cationic surfactants based on the alginic acid have the ability to decrease the interfacial tension between the light paraffin and water at 25 °C as it's depicted in Table S1 (supplementary). The surfactant with longer hydrocarbon chain length has lower interfacial tension. The interfacial tension of ALGOB, ALGDB and ALGHB aqueous solution are equal to 19, 17 and 14 mN/cm; respectively. The lower values of the interfacial tension, predict their ability to be used in enhancing oil recovery technology [65]. The synthesized polymeric surfactants based on the alginic acid have good emulsion

174

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

Table 6 Electrochemical parameters of impedance for of mild steel in 1.0 M HCl in the absence and presence of various concentrations from the synthesized polymeric cationic surfactants. Inhibitor name

Conc. of inhibitor (ppm)

Rs, (Ω·cm2)

Rct, (Ω·cm2)

Cdl, (F/cm2)

Ө

η%

Blank ALGOB

0.00 10 25 50 100 250 500 1000 10 25 50 100 250 500 1000 10 25 50 100 250 500 1000

1.307 1.81 2.29 1.1 2.41 1.9 2.26 2.069 2.26 2.43 1.85 0.72 2.24 2.05 2.19 2.40 2.87 2.70 3.59 3.29 3.60 2.98

155.9 283.8 310.3 345.7 425.8 503.8 526 610.2 306.9 344.9 392 505.4 548.3 616.8 723 342.8 409.8 475.8 556.4 650.1 806.4 933.2

129.42 70.76 64.60 58.04 47.08 44.23 42.36 41.19 65.32 58.12 51.14 44.07 40.62 36.11 34.77 58.48 54.35 46.81 40.03 38.67 35.23 30.44

00 0.4507 0.4976 0.5490 0.6339 0.6906 0.7036 0.7445 0.4920 0.5480 0.6023 0.6915 0.7157 0.7472 0.7844 0.5452 0.6196 0.6723 0.7198 0.7602 0.8067 0.8329

00 45.07 49.76 54.90 63.39 69.06 70.36 74.45 49.20 54.80 60.23 69.15 71.57 74.72 78.44 54.52 61.96 67.23 71.98 76.02 80.67 83.29

ALGDB

ALGHB

stability with paraffin oil, as it is clarified in the Table S1. The emulsion stability of the ALGOB, ALGDB and ALGHB aqueous solution with paraffin oil at 25 °C are equal to 3, 7 and 7 h, respectively. 3.5. Foaming power of the polymeric cationic surfactant The synthesized cationic polymeric surfactant characterized by their low foam power and stability, as it clear from the depicted experimental results in Table S1 (Supplementary). The foam height of the ALGOB, ALGDB and ALGHB polymeric surfactant are equal to 50, 15 and 10 mL, with foam stability equal to 1, 1.5 and 103 min, respectively. From the data, we can predict that the surfactant ALGOB could be used as corrosion inhibitors, washing machine laundry and as a biocide in oilfield sector, due to these applications need no foam [34]. 3.6. Biological activity Developing the antimicrobial materials is a vital case due to the fast upraising the microbial resistance toward the common antimicrobial. Surfactants were used as antimicrobial agents against most of bacteria and fungi trains. The activity of the synthesized ALGOB, ALGDB and ALGHB polymeric cationic surfactants against Gram positive (Bacillus subtilis and Staphylococcus aureus), Gram negative (Escherichia coli, Pseudomonas aeruginosa) and Fungi (Candida albicans, Aspergillus niger) have been performed using the disc diffusion method and the obtained results were depicted in Table 7. Analyzing these data, we reveal that the biological activity increases with increasing the hydrophobic

Inhibition zone diameter (mm/mg Sample) Bacteria species G (+ve)

4. Conclusion The three prepared polymeric cationic surfactants ALGOB, ALGDB and ALGHB based on the alginic acid have been prepared and elucidated its structures. The surface behavior of these surfactants in aqueous medium has been studied using surface tension and conductometric measurements, revealing that the surfactant ALGHB with higher hydrophobic chain length has lower critical micelle concentration as well as increasing temperature lead to increasing the micellization affinity. The synthesized polymeric surfactants characterized by their low foam power and stability. The ALGHB polymeric inhibitor with higher hydrocarbon chain length has higher carrion inhibition efficiency and it's increasing with raising the solution temperature up to 70 °C. The adsorption of the ALGOB, ALGDB and ALGHB polymeric inhibitors on the steel surface in 1.0 M HCl was found to obey the modified Langmuir isotherm. The Tafel curves revealed that theses surfactants are mixed type inhibitors. The prepared polymeric surfactants showed good antimicrobial effective against some bacteria and fungi, where the net magnitude of a surface parameter responsible for their antimicrobial effects difference and the surfactant ALGDB has an optimum antimicrobial effect. Appendix A. Supplementary data Supplementary data to this article can be found online at https://doi. org/10.1016/j.molliq.2018.10.017. References

Table 7 Antimicrobial activity of the synthesized cationic polymeric surfactant. Test organism

chain length then the antimicrobial activity decreases again [66–68]. The prepared polymeric cationic surfactants ALGOB, ALGDB and ALGHB have good biological activity against the tested bacteria and fungi as outlined in Table 7. The prepared surfactant ALGDB with twelve carbon atoms has the highest effect, where the diameter of inhibited zone is equal to 24, 20, 22, 24, 12 and 13 mm against the Bacillus subtilis (G+), Staphylococcus aureus (G+), Escherichia coli (G−), Pseudomonas aeruginosa (G−), Candida albicans and Aspergillus niger; respectively. The surface parameters of the synthesized polymeric cationic surfactants are responsible for their difference in the antimicrobial effect; where the magnitude effect of these parameters is responsible for their final effect. As described previously (Table 1), elongation the hydrophobic chain length of the synthesized polymeric cationic amphipathic, the critical micelle concentration decrease, consequently the maximum concentration of the surfactant at the cell wall decrease as it outlined in (Table 1). According to the activity is predicted to decrease with increasing the hydrophobic chain length. On the contrary; with increasing the hydrophobicity, the adsorption affinity at the interface is increase, consequently the biological activity is predicted to increase with increasing the surfactant chain length; i. e., the activity of ALGHB N ALGDB N ALGOB. Based on that, the magnitude of all these parameters is controlling their performance as an antimicrobial. According to the experimental data in (Table 7), we disclosed that the surfactant ALGDB is the most effective one against the tested bacteria and Fungi. The antibacterial performance of the prepared ALGOB, ALGDB and ALGHB polymeric surfactants against Bacillus subtilis were equal to 15, 24 and 20 mm, respectively (Table 7).

Fungi G (− ve)

Compuond ID

Bacillus subtilis

Staph. Aureus

Escherichia coli

Pseud. aeruginosa

Aspergillus niger

Candida albicans

ALGOB ALGDB ALGHB

15 24 20

15 20 17

13 22 18

15 24 17

7 12 10

8 13 11

[1] Eun-kyung Kang, Ga Young Jung, Seon Hwa Jung, Byung Min Lee, Synthesis and surface active properties of novel anionic surfactants with two short fluoroalkyl groups, J. Ind. Eng. Chem. 61 (2018) 216–226. [2] Ahmed H. Tantawy, Mohamed A. Abo-Riya, Samia M. Abdallah, Wagdy ElDougdoug, Novel cationic surfactants based on waste frying oil for cleaning water surface from petroleum films: synthesis, antimicrobial and surface properties, J. Mol. Liq. 253 (2018) 36–44. [3] Ismail Aiad, Mohamed M. EL-Sukkary, Ali El-Deeb, Moshira Y. El-Awady, Samy M. Shaban, Surface and biological activity of some prepared iminium surfactants based on Schiff bases, J. Surfactant Deterg. 16 (2013) 243–250. [4] Su Min Lee, JuYeon Lee, Hyon Pil Yu, Jong Choo Lim, Synthesis of environment friendly nonionic surfactants from sugar base and characterization of interfacial properties for detergent application, J. Ind. Eng. Chem. 38 (2016) 157–166.

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176 [5] Carla Serri, Bruno de Gennaro, Lilia Catalanotti, Piergiulio Cappelletti, Alessio Langella, Mariano Mercurio, Laura Mayol, Marco Biondi, Surfactant-modified phillipsite and chabazite as novel excipients for pharmaceutical applications, Microporous Mesoporous Mater. 224 (2016) 143–148. [6] Bernhard V.K.J. Schmidt, Valerio Molinari, Davide Esposito, Klaus Tauer, Markus Antonietti, Lignin-based polymeric surfactants for emulsion polymerization, Polymer 112 (2017) 418–426. [7] O. Kaczerewska, R. Leiva-Garcia, R. Akid, B. Brycki, Efficiency of cationic gemini surfactants with 3 azamethylpentamethylene spacer as corrosion inhibitors for stainless steel in hydrochloric acid, J. Mol. Liq. 247 (2017) 6–13. [8] Neetish Kumar Maurya, Ajay Mandal, Investigation of synergistic effect of nanoparticle and surfactant in macro emulsion based EOR application in oil reservoirs, Chem. Eng. Res. Des. 132 (2018) 370–384. [9] Rui Zhang, Jinhua Huo, Zhigang Peng, Qian Feng, Jian Zhang, Jixing Wang, Emulsification properties of comb-shaped trimeric nonionic surfactant for high temperature drilling fluids based on water in oil, Colloids Surf. A Physicochem. Eng. Asp. 520 (2017) 855–863. [10] E. Paluch, A. Piecuch, E. Obłąk, Ł. Lamch, K.A. Wilk, Antifungal activity of newly synthesized chemo degradable dicephalic-type cationic surfactants, Colloids Surf. B: Biointerfaces 164 (1) (2018) 34–41. [11] Samy M. Shaban, Ismail Aiad, Mohamed M. El-Sukkary, E.A. Soliman, Moshira Y. ElAwady, Preparation of capped silver nanoparticles using sunlight and cationic surfactants and their biological activity, Chin. Chem. Lett. 26 (2015) 1415–1420. [12] Anum Zahid, Aref Lashin, Usman Ali Rana, Nassir Al-Arifi, Imdad Ullah, Dionysios D. Dionysiou, Rumana Qureshi, Amir Waseem, Heinz-Bernhar Kraatz, Afzal Shah, Development of surfactant based electrochemical sensor for the trace level detection of mercury, Electrochim. Acta 192 (2016) 1007–1014. [13] Monika Srivastava, Preeti Tiwari, S.K. Srivastava, Ashish Kumar, Gopal Ji, Rajiv Prakasha, Low cost aqueous extract of Pisum sativum peels for inhibition of mild steel corrosion, J. Mol. Liq. 254 (2018) 357–368. [14] Lu Jiang, Yujie Qiang, Zulei Lei, Jianing Wang, Zhongjian Qin, Bin Xiang, Excellent corrosion inhibition performance of novel quinoline derivatives on mild steel in HCl media: experimental and computational investigations, J. Mol. Liq. 255 (2018) 53–63. [15] A. Jmiai, B. ElIbrahimi, A. Tara, S. ElIssami, O. Jbara, L. Bazzia, Alginate biopolymer as green corrosion inhibitor for copper in 1 M hydrochloric acid: experimental and theoretical approaches, J. Mol. Struct. 1157 (2018) 408–417. [16] Samy M. Shaban, A.S. Fouda, M.A. Elmorsi, T. Fayed, O. Azazy, A study on the effect of new prepared amide cationic amphipathic on the corrosion inhibition of API N80 steel pipelines in oil wells industries, Anti-Corr. Methods Mater. 65 (2) (2018) 197–209. [17] Kuen Yong Lee, David J. Mooney, Alginate: properties and biomedical applications, Prog. Polym. Sci. 37 (1) (2012) 106–126. [18] L. Zhia, Q. Li, Y. Li, Y. Song, Synthesis, adsorption and aggregation properties of new saccharide-cationic surfactants, Colloids Surf. A Physicochem. Eng. Asp. 436 (2013) 684–692. [19] Samy M. Shaban, Studying the effect of newly synthesized cationic surfactant on silver nanoparticles formation and their biological activity, J. Mol. Liq. 216 (2016) 137–145. [20] Olga Kaczerewska, Bogumił Brycki, Isabel Ribosa, Francesc Comelles, M. Teresa Garcia, Cationic gemini surfactants containing an O-substituted spacer and hydroxyethyl moiety in the polar heads: self-assembly, biodegradability and aquatic toxicity, J. Ind. Eng. Chem. 59 (2018) 141–148. [21] Khadidja Taleb, Mustapha Mohamed-Benkada, Nadjia Benhamed, Salima SaidiBesbes, Yves Grohens, Aicha Derdour, Benzene ring containing cationic gemini surfactants: synthesis, surface properties and antibacterial activity, J. Mol. Liq. 241 (2017) 81–90. [22] E.M.S. Azzam, M.F. Zaki, Surface and antibacterial activity of synthesized nonionic surfactant assembled on metal nanoparticles, Egypt. J. Pet. 25 (2) (2016) 153–159. [23] Ismail Aiad, Mohamed M. EL-Sukkary, Ali El-Deeb, Moshira Y. El-Awady, Samy M. Shaban, Surface properties, thermodynamic aspects and antimicrobial activity of some novel iminium surfactants, J. Surfactant Deterg. 15 (2012) 359–366. [24] Caili Dai, Sisi Fang, Mei Hu, Xiaojuan He, Ming wei Zhao, Xuepeng Wu, Shuai Yang, Yining Wu, Synthesis, surface adsorption and micelle formation of a class of morpholinium gemini surfactants, J. Ind. Eng. Chem. 54 (2017) 226–233. [25] L. Qi, Y. Fang, Z. Wang, N. Ma, L. Jiang, Y. Wang, Synthesis and physicochemical investigation of long alkyl chain betaine zwitterionic surfactant, J. Surfactant Deterg. 11 (2008) 55–59. [26] Ming Zhou, Gang Luo, Ze Zhang, Sisi Li, Chengwen Wang, Synthesis and properties evaluation of sulfobetaine surfactant with double hydroxyl, J. Mol. Struct. 1144 (2017) 199–205. [27] Divya Tripathy, V.K. Tyagi, Effect of fatty imidazolines on properties of laundry detergent compositions based on cationic surfactants, IOSR, J. Appl. Chem. 7 (5) (2014) 59–68. [28] Samy M. Shaban, Rabab M. El-Sherif, Mohamed A. Fahim, Studying the surface behavior of some prepared free hydroxyl cationic amphipathic compounds in aqueous solution and their biological activity, J. Mol. Liq. 252 (2018) 40–51. [29] ASTM standard designation G1-90: standard practice for preparing, cleaning, and evaluating corrosion test specimens, in: ASTM (Ed.), Annual Book of ASTM Standards, 03.02, American Society for Testing and Materials, Washington, D.C. 2000, pp. 15–21. [30] Samy M. Shaban, N (3 (dimethyl benzyl ammonio) propyl) alkanamide chloride derivatives as corrosion inhibitors for mild steel in 1M HCl solution: experimental and theoretical investigation, RSC Adv. 6 (2016) 39784–39800. [31] M.A. Hegazy, A novel Schiff base-based cationic gemini surfactants: synthesis and effect on corrosion inhibition of carbon steel in hydrochloric acid solution, Corros. Sci. 51 (2009) 2610–2618.

175

[32] O. Kaczerewska, R. Leiva-Garcia, R. Akid, B. Brycki, Efficiency of cationic gemini surfactants with 3 azamethylpentamethylene spacer as corrosion inhibitors for stainless steel in hydrochloric acid, J. Mol. Liq. 247 (2017) 6–13. [33] A.W. Bauer, W.M. Kirby, C. Sherris, M. Turck, Antibiotic susceptibility testing by a standardized single disk method, Am. J. Clin. Pathol. 45 (1966) 493–496. [34] Samy M. Shaban, A.S. Fouda, S.M. Rashwan, Hoyeda E. Ibrahim, M.F. El-Bhrawy, Synthesis and characterization of newly cationic surfactants based on 2 (2 (dimethylamino)ethoxy)ethanol: physiochemical, thermodynamic and evaluation as biocide, J. Mol. Liq. 221 (2016) 224–234. [35] M.T. Garcia, O. Kaczerewska, I. Ribosa, B. Brycki, P. Materna, M. Drgas, Hydrophilicity and flexibility of the spacer as critical parameters on the aggregation behavior of long alkyl chain cationic gemini surfactants in aqueous solution, J. Mol. Liq. 230 (2017) 453–460. [36] R. Kamboj, S. Singh, V. Chauhan, Synthesis, Colloids Surf. A Physicochem. Eng. Asp. 441 (2014) 233–241. [37] Samy M. Shaban, Ali A. Abd-Elaal, Studying the silver nanoparticles influence on thermodynamic behavior and antimicrobial activities of novel amide Gemini cationic surfactants, Mater. Sci. Eng. C 76 (2017) 871–885. [38] Caili Daia, Sisi Fang, Hu Mei, Xiaojuan He, Mingwei Zhao, Wu Xuepeng, Shuai Yang, Wu Yining, Synthesis, surface adsorption and micelle formation of a class of morpholinium gemini surfactants, J. Ind. Eng. Chem. 54 (2017) 226–233. [39] Samy M. Shaban, Ismail Aiad, Abdallah R. Ismail, Studying the surface parameters and determination the biological activity of N (3 (dimethyl benzyl ammonio) propyl) alkanamide chloride derivatives cationic surfactants, J. Surfactant Deterg. 19 (2016) 501–510. [40] Anwar Ali, Ummer Farooq, Sahar Uzair, Rajan Patel, Conductometric and tensiometric studies on the mixed micellar systems of surface-active ionic liquid and cationic surfactants in aqueous medium, J. Mol. Liq. 223 (2016) 589–602. [41] Samy M. Shaban, Ismail Aiad, Mohamed M. El-Sukkary, E.A. Soliman, Moshira Y. ElAwady, Synthesis of newly cationic surfactant based on dimethylamino propyl amine and their silver nanoparticles: characterization; surface activity and biological activity, Chin. Chem. Lett. 28 (2) (2017) 264–273. [42] Ting Zhang, Xianfei Cao, Xuchun Wang, Changchun Song, Synthesis, surface activity and thermodynamic properties of cationic gemini surfactants with diester and rigid spacers, J. Mol. Liq. 230 (2017) 505–510. [43] Darshak Bhatt, Kalpana Maheria, Jigisha Parikh, Mixed system of ionic liquid and non-ionic surfactants in aqueous media: surface and thermodynamic properties, J. Chem. Thermodyn. 74 (2014) 184–192. [44] Bo Gao, Mukul M. Sharma, A family of alkyl sulfate gemini surfactants. 1. Characterization of surface properties, J. Colloid Interface Sci. 404 (2013) 80–84. [45] Samy M. Shaban, A.S. Fouda, M.A. Elmorsi, T. Fayed, O. Azazy, Adsorption and micellization behavior of synthesized amidoamine cationic surfactants and their biological activity, J. Mol. Liq. 216 (2016) 284–292. [46] Nabel A. Negm, Ahmed F. El Farargy, Ahmad M. Al Sabagh, Naser R. Abdelrahman, New Schiff base cationic surfactants: surface and thermodynamic properties and applicability in bacterial growth and metal corrosion prevention, J. Surfactant Deterg. 14 (2011) 505–514. [47] N.A. Negm, N.G. Kandile, I.A. Aiad, M.A. Mohammad, New eco-friendly cationic surfactants: synthesis, characterization and applicability as corrosion inhibitors for carbon steel in 1 N HCl, Colloids Surf. A Physicochem. Eng. Asp. 391 (2011) 224–233. [48] Abass A. Olajire, Corrosion inhibition of offshore oil and gas production facilities using organic compound inhibitors - a review, J. Mol. Liq. 248 (2017) 775–808. [49] Yakun Zhu, Michael L. Free, Richard Woollam, William Durnie, A review of surfactants as corrosion inhibitors and associated modeling, Prog. Mater. Sci. 90 (2017) 159–223. [50] I.B. Obot, Ikenna B. Onyeachu, A. Madhan Kumar, Sodium alginate: a promising biopolymer for corrosion protection of API X60 high strength carbon steel in saline medium, Carbohydr. Polym. 178 (2017) 200–208. [51] Mohammed A. Amin, M.A. Ahmed, H.A. Arida, Taner Arslan, Murat Saracoglu, Fatma Kandemirli, Monitoring corrosion and corrosion control of iron in HCl by non-ionic surfactants of the TRITON-X series – part II. Temperature effect, activation energies and thermodynamics of adsorption, Corros. Sci. 53 (2011) 540–548. [52] T. Szauer, A. Brandt, Electrosorption bond theory in the Hartree–Fock approximation, Electrochim. Acta 26 (1981) 1209–1217. [53] Mohammad Mobin, Ruby Aslam, Jeenat Aslam, Non toxic biodegradable cationic gemini surfactants as novel corrosion inhibitor for mild steel in hydrochloric acid medium and synergistic effect of sodium salicylate: experimental and theoretical approach, Mater. Chem. Phys. 191 (2017) 151–167. [54] M.A. Bedaira, M.M.B. El-Sabbah, A.S. Fouda, H.M. Elaryian, Synthesis, electrochemical and quantum chemical studies of some prepared surfactants based on azodye and Schiff base as corrosion inhibitors for steel in acid medium, Corros. Sci. 128 (2017) 54–72. [55] Samy M. Shaban, Ismail Aiad, Mohamed M. El-Sukkary, E.A. Soliman, Moshira Y. ElAwady, Inhibition of mild steel corrosion in acidic medium by vanillin cationic surfactants, J. Mol. Liq. 203 (2015) 20–28. [56] R.F.V. Villamil, P. Corio, J.C. Rubim, S.M.L. Agostinho, Effect of sodium dodecyl sulfate on copper corrosion in sulfuric acid media in the absence and presence of benzotriazole, J. Electroanal. Chem. 472 (2) (1999) 112–119. [57] Ismail Aiad, Samy M. Shaban, Mohamed M. El-Sukkary, Moshira Y. El-Awady, E.A. Soliman, Electrical and gravimetric estimation the corrosion inhibition of three synthesized cationic surfactants N (3 (butylidene amino) propyl) N,N dimethyl alkan 1 ammomium bromide derivatives in 1 M HCl, J. Mater. Perform. Charact. 6 (1) (2017) 429–450. [58] I. Aiad, Samy M. Shaban, H.Y. Moustafa, Ahmed Hamed, Experimental investigation of newly synthesized Gemini cationic surfactants as corrosion inhibitors of mild steel in 1.0 M HCl, Prot. Met. Phys. Chem. Surf. 54 (2018) 135–147.

176

S.M. Shaban et al. / Journal of Molecular Liquids 273 (2019) 164–176

[59] S. Abd-ElWanees, N.M. ElBasiony, A.M. Al-Sabagh, M.A. Alsharif, S.M. Abd El Haleem, M.A. Migahed, Controlling of H2 gas production during Zn dissolution in HCl solutions, J. Mol. Liq. 248 (2017) 943–952. [60] K. Zakaria, N.A. Negm, E.A. Khamis, E.A. Badr, Electrochemical and quantum chemical studies on carbon steel corrosion protection in 1 M H2SO4 using new ecofriendly Schiff base metal complexes, J. Taiwan Inst. Chem. Eng. 61 (2016) 316–332. [61] Turuvekere K. Chaitra, Kikkeri N. Mohana, Doddahosuru M. Gurudatt, Harmesh C. Tandon, Inhibition activity of new thiazolehydrazones towards mild steel corrosion in acid media by thermodynamic, electrochemical and quantum chemical methods, J. Taiwan Inst. Chem. Eng. 67 (2016) 521–531. [62] A. Bahgat Radwan, Mostafa H. Sliem, Paul C. Okonkwo, Mohamed F. Shibl, Aboubakr M. Abdullah, Corrosion inhibition of API X120 steel in a highly aggressive medium using stearamidopropyl dimethylamine, J. Mol. Liq. 236 (2017) 220–231. [63] Hany M. Abd El-Lateef, Ahmed M. Abu-Dief, Mounir A.A. Mohamed, Corrosion inhibition of carbon steel pipelines by some novel Schiff base compounds during acidizing treatment of oil wells studied by electrochemical and quantum chemical methods, J. Mol. Struct. 1130 (2017) 522–542.

[64] K. Zakaria, A. Hamdy, M.A. Abbas, O.M. Abo-Elenien, New organic compounds based on siloxane moiety as corrosion inhibitors for carbon steel in HCl solution: weight loss, electrochemical and surface studies, J. Taiwan Inst. Chem. Eng. 65 (2016) 530–543. [65] N.A. Negm, A.S. El-Tabl, I.A. Aiad, K. Zakareya, A.H. Moustafa, J. Surfactant Deterg. 16 (2013) 857–863. [66] J. Haldar, P. Kondaiah, S. Bhattacharya, Synthesis and antibacterial properties of novel hydrolyzable cationic amphiphiles incorporation of multiple headgroups leads to impressive antibacterial activity, J. Med. Chem. 48 (2005) 3823–3831. [67] Ali A. Abd-Elaal, N.M. Elbasiony, Samy M. Shaban, E.G. Zaki, Studying the corrosion inhibition of some prepared nonionic surfactants based on 3 (4 hydroxyphenyl) propanoic acid and estimating the influence of silver nanoparticles on the surface parameters, J. Mol. Liq. 249 (2018) 304–317. [68] M.C. Murguia, V.A. Vaillard, V.G. Sanchez, G. De Conza, R.J. Grau, Synthesis, surfaceactive properties and antimicrobial activities of new double-chain gemini surfactants, J Oleo Sci 57 (5) (2008) 301–308.