Accepted Manuscript Title: Synthesis of Mesoporous MgO Nanostructures Using Mixed Surfactants Template for Enhanced Adsorption and Antimicrobial activity Authors: Jyoti Sharma, Manisha Sharma, Soumen Basu PII: DOI: Reference:
S2213-3437(17)30319-6 http://dx.doi.org/doi:10.1016/j.jece.2017.07.015 JECE 1730
To appear in: Received date: Revised date: Accepted date:
17-4-2017 3-7-2017 5-7-2017
Please cite this article as: Jyoti Sharma, Manisha Sharma, Soumen Basu, Synthesis of Mesoporous MgO Nanostructures Using Mixed Surfactants Template for Enhanced Adsorption and Antimicrobial activity, Journal of Environmental Chemical Engineeringhttp://dx.doi.org/10.1016/j.jece.2017.07.015 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Synthesis of Mesoporous MgO Nanostructures Using Mixed Surfactants Template for Enhanced Adsorption and Antimicrobial activity
Jyoti Sharma, Manisha Sharma and Soumen Basu*
School of Chemistry and Biochemistry, Thapar University, Patiala 147004, India. E-mail:
[email protected]. Abstract: Mesoporous MgO nanostructures with the high specific surface area (180 m2g-1) were synthesized via mixed surfactant-templating method. The synthesis of MgO with mixed surfactants system has rarely been discussed in the literature, which motivates us to synthesize MgO in this system. Alteration in catanionic surfactant molar ratio and chain length resulted in a change in surface area. Morphological and physical properties were analyzed using FESEM, HRTEM, FTIR and XRD. Synthesized MgO nanorods showed excellent adsorption properties for dye removal from aqueous solution. The obtained adsorption capacities were 333.33, 250 and 200 mg.g -1 for Methylene Blue (MB), Alizarin red (AZ) and Rhodamine B (RD), respectively, which is higher or comparable with other reported methods in literature. Other than adsorption, MgO can act as a good bactericide. The effective antimicrobial activity of MgO was analyzed via minimal inhibitory concentration (MIC) method against E. coli and B. subtilis and obtained IC50 (half maximal inhibitory concentration) values after 24h were 71.98±0.03 and 94.01±0.030 respectively. Keywords: MgO nanostructures; Adsorption; Dyes; Mesoporous; Antimicrobial activity
1
1. Introduction:
Currently, research based on the nanomaterials are of more interest due to their unique physicochemical properties [1]. The combination of nanomaterials and porous structures represents the most interesting of the rapidly growing areas.
Porous material includes
microporous (<2 nm), mesoporous (2-50 nm) and macroporous (>50 nm) which have interconnected pore network of solid composites. Other than this, natural substances such as clays, biological tissues (e.g. bones), rocks, and synthetic materials together with metal oxides, ceramics, membranes and carbonaceous materials can be considered as porous material [2]. Porous solid catalysts with high surface area and different pore size control the adsorption of various toxic organic molecules [3-7]. Out of porous materials, mesoporous materials have been of great interest to the researchers. Metal oxide nanostructures are given more and more importance because of their high chemical activity and their specificity in interaction [8]. Many researchers have submitted their reports on the synthesis of mesoporous materials like ZnO [9], TiO2 [10-13], CaO [14], MgO etc [15]. MgO is a solid of highly ionic nature, crystalline structure and have simple stoichiometry which can be widely prepared in different shapes and sizes. Adsorption properties are affected by the morphology MgO crystals [16]. Earlier MgO was prepared by sol-gel method [17, 18] and thermal decomposition of magnesium carbonate and hydroxides [19, 20]. Many methods including the thermal oxidation [21], thermal reduction [22] and hydrophobic interactions (which helps in self-assembly of building blocks and templateassisted synthesis) have been used to prepare metal oxide nanostructures [23]. [24] To overcome the problem of agglomeration, MgO is also obtained from the preferable ‘surfactant templating method’ over Sol-gel method [25]. Surfactants play a significant role in the synthesis of mesoporous structures or composites. Their behaviour can be determined by the nature of bonding between counter ions. The ionogenic Cetyltrimethylammonium bromide (CTAB) act as a soft template [26]. The long alkyl chains of surfactants help in forming the small voids in the 2
composite material [27]. Other than the concentrations the arrangements of the templates depends upon the chemical nature, sizes and charges of the micelles formed from the surfactant.The arrangement of these templates depends on the concentrations, chemical nature of the used surfactants, their sizes and charges of the micelles formed [28]. Therefore, different surfactant templates may result in different structures and morphologies of inorganic materials. MgO with surfactant templating method results in formation of the rod-like structure [29]. Selfassembly and self-arrangement of surfactants helps in making the internal hollow sphere of MgO [30]. A cationic surfactant (CTAB) had been used in the sol-gel reaction, in which only spherical MgO was obtained, with less surface area [31]. MgO has been synthesized by using the mixtures of cationic and anionic surfactants in molar ratios from which 220 m2g-1 surface area was recorded [32]. The literature based on the synthesis of high surface area MgO nanomaterials using mixed surfactant systems has been rarely investigated. The adsorption of pollutants on metal oxides is an interesting topic for both academic and industrial fields. Organic dyes are carcinogenic and acts as an obstacle in reusing the wastewater for various purposes. To maintain the human health, to preserve the aquatic life and to create an overall balance in the universe, a highly economical, efficient and a non-toxic method should be devised for removal of dyes [33]. Adsorption is a cost effective process which helps in removing synthetic dyes and hazardous pollutants from wastewater. Ahmed et al., in 2016 discussed the adsorption of fast orange and bromophenol blue dyes on MgO and reported adsorption capacities of 30 and 40 mg.g−1 respectively [34]. Hu et al., 2010 discussed the adsorption of X3B and congo red on MgO nanosheets and calculated the adsorption capacities of 303.3 and 278.5 mg.g1 respectively [35]. MgO also has an excellent bactericidal property against the different genre of bacterias. Inorganic antimicrobial agents proved to be more safe and stable under high temperatures. For a good bactericide, high surface area metal oxide is required. Powder metal oxides, ZnO [36], CaO [37], MgO
3
[38] showing good antibacterial activity against S. aureus and E. coli has been evaluated by some researchers. Nanorods have much higher efficiency as bactericide due to their high surface area. The formation of reactive oxygen species such as superoxide, explained the antibacterial behaviour of MgO. It had also been reported that increase in the surface area of MgO with the increase in concentration of superoxide in solution, results into more effective destruction of the bacterial cell wall[39]. In this study, the formation mechanism of MgO nanorods synthesized via mix surfactants templating method has been discussed in detail. Also explored here are the effects of different molar concentration of catanionic surfactants and alteration in their chain length on the surface properties of synthesized MgO nanorods. MgO nanorods with high surface area are used as an efficient adsorbent for organic dyes like MB, AZ, RD from aqueous solutions. Kinetic and thermodynamic equilibrium studies were performed to measure the adsorption efficiency for different dyes. Other than adsorption, antimicrobial activity of MgO was analyzed via MIC method against E. coli and B. subtilis. 2. Experimental section 2.1 Material and methods
All reagents used for synthesis of MgO were of analytical grade and all solutions were prepared in Distilled water. Magnesium chloride, dodecyltrimethylammonium bromide (C12TAB), tetradecyltrimethylammonium bromide (C14TAB), cetyltrimethylammonium bromide(C16TAB) octadecyltrimethylammonium bromide (C18TAB), sodium hydroxide (NaOH), methylene blue (MB), sodiumdodecyl sulfate (SDS), alizarin red (AZ) and rhodamine B (RD) were purchased from Merck, India. Magnesium chloride was used as a magnesium source. An aqueous solution of MgCl2 (1M) was added with different molar ratio of C16TAB+SDS mixture and the solution was stirred for 1h after which 2M of NaOH solution was added drop-wise and stirred continuously for another 30 minutes. Finally, the white precipitate of Mg(OH)2 was obtained by centrifugation (5000 rpm, 5 min), followed by washing (5
4
times) with distilled water and dried up at 80˚C for 12h. Calcination was done at 400˚C with a heating ramp of 1̊ C/min. The different ratio of C16TAB+SDS mixture used for the synthesis of MgO is given below. For a variation of chain length of CnTAB (n=12-18), we have followed the procedure mentioned above keeping the equimolar concentration of CnTAB: SDS = 0.125: 0.125. For comparison purposes, we have also synthesized blank MgO by following the same procedure with the absence of surfactants and mentioned it as a blank sample.
2.2 Characterization: To verify the crystal structure of mesoporous MgO powder, X-ray diffraction measurements was done at room temperature using Pan Analytical (X’Pert-pro) diffractometer with Cu Kα1 radiation (λ= 1.5406 Å) with a scanning interval (2Ɵ) from 15̊ to 75̊. Brunauer Emmett Teller (BET) was used to determine the specific pore sizes and surface areas using BEL mini-II, Micro Trac Corp. Pvt. Ltd, Japan. Degassing of samples was done at 200 °C in vacuum for more than 2 h prior to the measurements. Fourier Transform Infrared Spectroscopy (FTIR) of the samples was obtained by using Agilent Technologies, Carry 660 FTIR spectrophotometer. Field Emission Scanning Electron Microscopic (FESEM) JEOL JSM-6510LV coupled with Energy dispersive spectroscopy (EDS) was used to detect the morphology and elemental components of MgO nanostructures. Transmission Electron Spectroscopy (TEM) images were obtained with FEI Technai G2 F20 operating at 200KV. 2.3 Adsorption study: Adsorption of dyes on mesoporous MgO was performed in a batch system at ambient temperature using favorable conditions. Each dye underwent to the same procedure which is discussed below. Mesoporous MgO (0.01g) was added into different dyes solution (100-400 mgL-1) and agitated for different time periods at a speed of 200 rpm. After the adsorption procedure, the solutions were centrifuged for 10 min at 5000 rpm. After centrifugation, the samples were analyzed by UV–Visible spectrophotometer (Champion UV-500) by monitoring the absorbance changes at λmax 660 nm for MB,
5
600 nm for AZ and 510 nm for RD. The adsorption capacity (qe) was calculated by the uptake amount of dye adsorbed per unit mass of mesoporous MgO (mg/g) using the formula:
qe
( Co Ce ) V m
(1)
Where Co is the initial and Ce is the equilibrium concentrations (mg/L), while V is the volume of the solution (L) and m is the weight of the adsorbent (g). 2.4 Antibacterial activity for Mesoporous MgO: The antimicrobial activity of mesoporous MgO nanomaterials against E. coli (MTCC-77) and B. subtilis (MTCC-441) was analyzed by using the MIC method. Bacterial cells were incubated with various concentrations (0 to 600 µg) of MgO. After incubation, optical density was analyzed at 600 nm with MgO for a period of 24 h at 35 ̊C and results for the mean value of 3 mutually independent experiments were plotted. 3. Results and discussion: This work clearly shows the synthesis and characterization of MgO by using a mix surfactant system. Effect of different chain lengths and different ratios of surfactants alter the porosity as well as the surface area of synthesized nanomaterial, which was analyzed by BET surface area analyzer. 3.1 Effect of different molar ratios of surfactants on MgO The surfactant can decrease the surface energy and prevent agglomeration when adsorbed on the surface of the nanomaterials. Polymeric dispersants stabilize and control the size of the nucleating particles [40]. To check the effect of mixed surfactant on surface area, a mixture (1:1) of cationic (CTAB) and anionic (SDS) surfactants were used. Here, mesostructured MgO nanorods were synthesized using equimolar composition (1:1) of C16TAB and SDS (0.125, 0.25, 0.5, 1, 1.5 M) at ambient temperature. Fig. 1a represents the N2 adsorption- desorption curves for the mesoporous MgO. According to BrunauerEmmett-Teller (BET) model and IUPAC classification, all the curves are of type IV with H3 hysteresis loop
6
and the loop increases between 0.8
7
nanorods are formed [44]. Fig. 3 describes the mechanism of the nucleation and formation of MgO nanorods. 3.2 Effect of chain length of surfactant: To check the effect of chain length on the surface structure of MgO, we have varied the chain length of only cationic surfactant, CnTAB (n=12-18) keeping the molar ratio of CTAB and SDS fixed to 0.125M. The chain length is an important factor to change the pore size of the compound as it alters the size of the resulting micelle. The increment in the alkyl groups (n=12 to 18) of cationic surfactant leads to the increase in the surface area. Fig. 4(a-b) shows the nitrogen adsorption-desorption and pore volume distribution curve using differential BJH plot for the synthesized MgO nanorods with a different chain length of CTAB. From the IUPAC classification, all are type IV and exhibit the characteristic hysteresis loop of mesoporous materials. Fig. 4(a-b) clearly shows the increase in surface area (50 to 180 cm2g-1) and decrease in pore size (40-19nm) with the increasing chain length of CnTAB (n=12 to 18). Steric hindrance of the long alkyl chains of cationic surfactants results in the decrease in the pores size due to the agglomeration of the compound. Detailed N2 adsorption-desorption parameters for synthesized MgO nanorods with a different chain length of CTAB are shown in Table 3. The morphology of MgO nanorods with a different chain length of CTAB was analyzed by FESEM. Fig. 5(a-b) shows the appearance of long (150-200 nm) rod-shaped MgO with some irregular structures. HRTEM analysis Fig. 5(c-d) also confirms the rod shape structure of MgO. Although the surface properties of different MgO synthesized by varying the chain length of surfactant is different, but the morphology (size and shape) of the MgO is almost same (diameters of all nanorods are almost similar). The elemental composition of the MgO was confirmed by the EDS analysis as shown in Fig. 5e which confirms the purity of MgO. The crystalline nature of the mesoporous MgO nanorods with a different chain length of CTAB was analyzed through their XRD pattern. Fig. 6a presents the X-ray diffraction pattern of the MgO nanorods by using a mixture of Cn=12-18TAB+SDS and blank MgO (without surfactant). The XRD pattern
8
show the Bragg’s reflections at 2θ = 37.8̊, 42.5̊ and 63.54̊ indexed to (111), (200), (220) planes respectively, confirming crystalline nature of MgO by comparison with the JCPDS card no. 45-0946. Fig. 6b shows the FTIR spectra of the synthesized mesoporous MgO nanorods and the blank MgO. The broad absorption band at 578cm-1 confirms the symmetric stretching of Mg=O bond. The absorption bands at 3345 cm-1 (stretching) and 1431.7 cm-1 (bending) indicate the presence of hydroxyl groups (OH), which is due to the presence of Mg(OH)2. The bands at 1095cm-1 are due to C-O stretching. 3.3 Adsorption study: 3.3.1 Effect of pH pH is a significant parameter which affects both surface binding sites and aqueous chemistry of the adsorbents. To check the effect of pH on the adsorption process we have changed the pH of the solution from 3-10 by adding 0.1 N NaOH and 0.1 N HNO3 solution. Fig. 7a shows the results of the effect of pH on MB, AZ, RD removal efficiencies. The maximum adsorption for all the three dyes was noted at pH6 which clearly signifies the presence of maximum binding sites. The adsorption of MB, AZ, RD using MgO increases from 72.4 to 96.8 %; 71.2 to 95.4% and 69.8 to 94.2% respectively, by adjusting the pH. At lower pH, adsorbent’s surface was surrounded by hydronium ions that compete with the dye, which prohibit the dye from approaching the binding sites on the adsorbents [45, 46]. So with the increase in pH, there is less competition between the hydronium ions and the adsorbent’s surface, which enhances the adsorption. 3.3.2 Effect of contact time Effect of contact time was analyzed between 10 to 120 min, which elaborated the efficiency of MgO on dyes. At 6 pH, the initial concentration of all the three dyes (MB, AZ, RD) was 100 mg/L. Fig. 7b shows that with increase in time, there is an increase in the adsorption capacity (qe). It also indicates that MB has the maximum absorbance with time in comparison to AZ and RD. The kinetics data explained that there is the presence of two phases: (i) fast phase shows the maximum adsorption and (ii) saturation
9
phase explains the phase of equilibrium [47, 48].There is no significant increase in the adsorption after 90 min for all the three dyes. 3.3.3 Effects of dye concentration: As the concentration of dyes (MB, AZ, RD) affects the adsorption capacity of mesoporous MgO (0.01g), the adsorption studies were done with different concentration (100-400 mg/L) of dyes. Fig. 7c shows the increase in adsorption capacity with the increase in the concentration of dyes which results to increase in the mass gradient between the solution and the adsorbent. Dye adsorption experiments on the MgO were performed in a batch system at room temperature using favorable conditions (pH-6). 3.3.4 Effect of temperature on adsorption: To determine the effect of temperature on dyes (MB, AZ, RD), MgO was treated at 293, 313, 333 and 353 K respectively. The observed decrease in values of 1/T and ln(qe/Ce) with high temperature indicates the endothermic nature of the adsorption process [49]. The values of ln(qe/Ce) at different temperatures were analyzed according to Van’t Hoff equation:
q H S ln e RT R Ce
(2)
Where R represents the universal gas constant (8.314 J/(mol K)) and T defines the absolute temperature (in Kelvin). A linear graph was obtained by plotting ln(qe/Ce) against 1/T Fig. 7d. Endothermic nature of adsorption process was confirmed by the positive value of ∆H and it states that the adsorption is favored at a higher temperature. Gibbs free energy of adsorption (∆G) is calculated from the following relation:
G H TS
(3)
Results from Table 4 show the adsorption reaction were spontaneous and confirmed by the negative values of ∆G.
10
3.3.5 Kinetic study: The kinetics of MB, AZ, RD adsorption on MgO were also evaluated using Lagergren’s first-order rate equation and pseudo-second-order kinetic model. Lagergren’s first-order rate equation is usually used to calculate adsorption kinetic and expressed as;
log(qe qt ) log(qe )
K1 t 2.303
(4)
The pseudo-second-order rate equation is expressed as;
t 1 t qt k2 qe2 qe
(5)
where qt and qe are the amount of metal oxide (mg/g) and a dye adsorbed at equilibrium and at any given time (min) respectively. While K1 and k2 represent the rate constant for the pseudo-first order reaction for adsorption (min-1) and pseudo-second order reaction (g/mg min) for adsorption respectively [50]. Fig. 8(a-b) shows the linear fit plot for pseudo-first and pseudo-second-order rate equation. Table 5 shows the kinetic parameters obtained from pseudo-first and pseudo-second-order kinetic model. In all cases, correlation coefficients were found to be closer to unity, but the proposed results suggest pseudo-second-order is best fitted in comparison to pseudo-first for adsorption due to the chemisorption process. (# : mg.g-1, @ : min-1, + : g/mg min) 3.3.6. Intraparticle Diffusion Weber and Morris described that diffusion of the adsorbate from the surface to the internal pores of the adsorbent and referred it as intaparticle diffusion. The intraparticle diffusion model was analysed using kinetic study and mathematical relation is as follows:
qt K diff t 1/ 2 C
11
where qt (mg g-1) is the amount of dye adsorbed at time t and Kdiff (mg g-1 min-1/2) is the rate constant for intraparticle diffusion. The thickness of the boundary layer can be calculated from the value of C, large intercept describes the great boundary layer effect. A plot of qt versus t1/2 can give a linear or multi linear and proves the involvement of intraparticle diffusion in the adsorption process. The linear graph from the origin suggests the intaparticle diffusion as a rate limiting step. The slowest step i.e, the rate limiting step control the overall adsorption process. The intraparticle diffusion can be explained by two parts, due to external diffusion the starting part of the adsorption is rapid and latter part said to be the rate limiting step known as intra-particle mass transfer rate [51]. Therefore, intraparticle diffusion is one of the factors which affect the rate of ease for the adsorption to reach equilibrium state. Fig 8c illustrates Weber and Morris plot for all the three dyes MB, AZ, RD adsorption on MgO. Due to multi-linearity correlation it explained the significant role in adsorption mechanism. The graph indicates that it is not the solely rate limiting step. 3.3.7 Adsorption isotherms: Data obtained from adsorption isotherm played a significant role in establishing the effectiveness of adsorption. The association between the concentrations of adsorbed and dissolved adsorbate at equilibrium along with the interactive behavior between the adsorbate and adsorbent can also be described by adsorption isotherm. In this study, Langmuir and Freundlich isotherm models were used to analyze the adsorption mechanisms, through which experimental results of dye can be explained in a wide variety of concentrations. The Langmuir equation is expressed as follows;
Ce C 1 e qe Qob Qo
(6)
where the equilibrium adsorption capacity is qe (mg/g) and the maximum amount of the MB, AZ and RD adsorbed per unit weight of the adsorbent is Q0 (mg/g). The linear plots of Ce/qe vs. Ce propose the validity of the Langmuir isotherms and the values of Q0 and b were obtained from slope and intercepts
12
of the plots. The calculated maximum adsorption capacity of MgO nanorods for MB, AZ and RD were 333.3, 250 and 200 mg/g respectively. Also, when the surface is fully covered with dye, Q0 represents adsorption capacity, helping in the assessment of adsorption action of adsorbents. Langmuir equilibrium constant (b) which is related to the connecting spots and explains the bond energy for the adsorption reaction [52]. The Freundlich isotherm is a model of multilayer adsorption on the adsorbent which can be described as:
log qe log K f
log Ce n
(7)
Freundlich constants, Kf (mg1-1/n L1/n g-1) and n depict the adsorption capacity and intensity, respectively[53]. Fig. 9a and Fig. 9b showed the plots for Langmuir isotherm and Freundlich model for the adsorption of dye on MgO. Table 6 shows the fitted parameters (constants and correlation coefficients) from experimental data for both Freundlich and Langmuir isotherms. (where, * : mg1-1/nL1/ng-1, # : mg.g-1) 3.3.9. Error analysis: The non-linear regression was calculated using the software IBM SPSS Statistics 20 for the determination of isotherm models. Only R2 (coefficient of determination) is not only the non linear analysis which is important, there is one more non-linear analysis to evaluate the goodness of the models and describe the sorption phenomena. The correlation of the experimental data can further be estimated with the help of chi-square test and is given as:
2
(qe ,exp qe,cal ) 2 qe,cal
Where qe ,exp and qe,cal are the experimental and model equilibrium capacity data, respectively. if the data from the model are similar to the experimental data , the value of 2 will be small, otherwise it will be large.
13
3.3.10. Desorption Studies: Synthesised MgO nanorods were found to be reusable for atleast 5 cycles after adsorption study for toxic dyes. After completion on each adsorption cycle, adsorbates were easily separated from aqueous solution of toxic pollutants by simple filtration or centrifugation process. No additional filtration techniques were required, so these adsorbates are considered as user-friendly and cost effective. Fig.10 shows the plot of percentage removal vs. no. of cycles. 3.3.10. Antimicrobial activities: The antibacterial activity of mesoporous MgO against E. coli and B. Subtilis is shown in Fig. 11. Being proportional to the concentration of the nanorods used, the antibacterial activity of MgO was analyzed via MIC method against E. coli and B. subtilis respectively. Obtained were the IC50 values i.e, the half maximal inhibitory concentration after 24h were 71.98±0.03 and 94.01±0.030, respectively. As the pore size of the mesoporous MgO decreases with the increase in specific surface area, the potential number of reactive groups on the particle surface increases, due to which high antibacterial activity was expected.[54] Conclusion: In summary, mesoporous MgO nanorods were synthesized by mixed surfactant templating method and analyzed were the effect of surfactants ratio and chain length on pore size distribution, surface area, and morphology. At particular concentration (0.125M) of mixed cat-anionic surfactant, the surface area becomes high (180 m2g-1) may be due to generation of the lamellar crystallinity for mesostructured MgO. Adsorption study showed that pseudo second order kinetic and Langmuir isotherm models were best fitted as values of correlation coefficients were more near to unity. The effects of adsorbate pH, temperature, contact time and adsorbent concentration were also explored. The antibacterial activity of mesoporous MgO nanorods was investigated against the gram-negative E. coli and gram-positive B. subtilis bacteria by MIC method.
14
Conflicts of interest: The authors declare that there is no conflict of interests regarding the publication of this paper. Acknowledgement: The authors are grateful to BRNS (Grant no 34/14/63/2014) for financial assistance and Thapar University (seed money grant) for infrastructure/instrumental facilities. References: [1] Z. Zhang, Q. Fu, Y. Xue, Z. Cui, S. Wang, Theoretical and Experimental Researches of Size-Dependent Surface Thermodynamic Properties of Nanovaterite, The Journal of Physical Chemistry C, 120 (2016) 21652-21658. [2] R.J. White, R. Luque, V.L. Budarin, J.H. Clark, D.J. Macquarrie, Supported metal nanoparticles on porous materials. Methods and applications, Chemical Society Reviews, 38 (2009) 481-494. [3] B. Neppolian, Q. Wang, H. Jung, H. Choi, Ultrasonic-assisted sol-gel method of preparation of TiO 2 nano-particles: characterization, properties and 4-chlorophenol removal application, Ultrasonics sonochemistry, 15 (2008) 649-658. [4] K. Mageshwari, S.S. Mali, R. Sathyamoorthy, P.S. Patil, Template-free synthesis of MgO nanoparticles for effective photocatalytic applications, Powder technology, 249 (2013) 456-462. [5] M. Rezaei, M. Khajenoori, B. Nematollahi, Synthesis of high surface area nanocrystalline MgO by pluronic P123 triblock copolymer surfactant, Powder technology, 205 (2011) 112-116. [6] H. Niu, Q. Yang, K. Tang, Y. Xie, Self-assembly of porous MgO nanoparticles into coral-like microcrystals, Scripta materialia, 54 (2006) 1791-1796. [7] B. Nagappa, G. Chandrappa, Mesoporous nanocrystalline magnesium oxide for environmental remediation, Microporous and Mesoporous Materials, 106 (2007) 212-218. [8] M. Boutonnet, A. Marinas, V. Montes, R. Suárez-Paris, M. Sánchez-Domınguez, Nanocatalysts: Synthesis in Nanostructured Liquid Media and Their Application in Energy and Production of Chemicals, in: Nanocolloids: A Meeting Point for Scientists and Technologists, Elsevier, 2016, pp. 211. [9] M. Montazer, M. Maali Amiri, ZnO nano reactor on textiles and polymers: ex situ and in situ synthesis, application, and characterization, The Journal of Physical Chemistry B, 118 (2014) 1453-1470. [10] J.C. Yu, L. Zhang, Z. Zheng, J. Zhao, Synthesis and characterization of phosphated mesoporous titanium dioxide with high photocatalytic activity, Chemistry of Materials, 15 (2003) 2280-2286. [11] A. Mehta, M. Sharma, A. Kumar, S. Basu, Effect of Au content on the enhanced photocatalytic efficiency of mesoporous Au/TiO2 nanocomposites in UV and sunlight, Gold Bulletin, 1-9. [12] A. Mishra, A. Mehta, M. Sharma, S. Basu, Enhanced heterogeneous photodegradation of VOC and dye using microwave synthesized TiO 2/Clay nanocomposites: A comparison study of different type of clays, Journal of Alloys and Compounds, 694 (2017) 574-580.
15
[13] A. Mishra, A. Mehta, M. Sharma, S. Basu, Impact of Ag nanoparticles on photomineralization of chlorobenzene by TiO 2/bentonite nanocomposite, Journal of Environmental Chemical Engineering, 5 (2017) 644-651. [14] X. Li, L. Zhang, X. Dong, J. Liang, J. Shi, Preparation of mesoporous calcium doped silica spheres with narrow size dispersion and their drug loading and degradation behavior, Microporous and mesoporous materials, 102 (2007) 151-158. [15] M. Bhagiyalakshmi, J.Y. Lee, H.T. Jang, Synthesis of mesoporous magnesium oxide: its application to CO 2 chemisorption, International Journal of Greenhouse Gas Control, 4 (2010) 51-56. [16] S.-W. Bian, Z. Ma, Z.-M. Cui, L.-S. Zhang, F. Niu, W.-G. Song*, Synthesis of Micrometer-Sized Nanostructured Magnesium Oxide and Its High Catalytic Activity in the Claisen− Schmidt Condensation Reaction, The Journal of Physical Chemistry C, 112 (2008) 15602-15602. [17] N.K. Nga, P.T.T. Hong, T. Dai Lam, T.Q. Huy, A facile synthesis of nanostructured magnesium oxide particles for enhanced adsorption performance in reactive blue 19 removal, Journal of colloid and interface science, 398 (2013) 210-216. [18] Z. Wu, C. Xu, H. Chen, Y. Wu, H. Yu, Y. Ye, F. Gao, Mesoporous MgO nanosheets: 1, 6-hexanediaminassisted synthesis and their applications on electrochemical detection of toxic metal ions, Journal of Physics and Chemistry of Solids, 74 (2013) 1032-1038. [19] R. Sathyamoorthy, K. Mageshwari, S.S. Mali, S. Priyadharshini, P.S. Patil, Effect of organic capping agent on the photocatalytic activity of MgO nanoflakes obtained by thermal decomposition route, Ceramics International, 39 (2013) 323-330. [20] G. Yuan, J. Zheng, C. Lin, X. Chang, H. Jiang, Electrosynthesis and catalytic properties of magnesium oxide nanocrystals with porous structures, Materials Chemistry and Physics, 130 (2011) 387-391. [21] J.D. Majumdar, U. Bhattacharyya, A. Biswas, I. Manna, Studies on thermal oxidation of Mg-alloy (AZ91) for improving corrosion and wear resistance, Surface and Coatings Technology, 202 (2008) 36383642. [22] H.J. Jeong, K.H. An, S.C. Lim, M.-S. Park, J.-S. Chang, S.-E. Park, S.J. Eum, C.W. Yang, C.-Y. Park, Y.H. Lee, Narrow diameter distribution of singlewalled carbon nanotubes grown on Ni–MgO by thermal chemical vapor deposition, Chemical physics letters, 380 (2003) 263-268. [23] J. Lv, L. Qiu, B. Qu, Controlled growth of three morphological structures of magnesium hydroxide nanoparticles by wet precipitation method, Journal of Crystal Growth, 267 (2004) 676-684. [24] K. Sanosh, A. Balakrishnan, L. Francis, T. Kim, Sol–gel synthesis of forsterite nanopowders with narrow particle size distribution, Journal of Alloys and Compounds, 495 (2010) 113-115. [25] W.-C. Li, A.-H. Lu, C. Weidenthaler, F. Schüth, Hard-templating pathway to create mesoporous magnesium oxide, Chemistry of materials, 16 (2004) 5676-5681. [26] X. Li, W. Xiao, G. He, W. Zheng, N. Yu, M. Tan, Pore size and surface area control of MgO nanostructures using a surfactant-templated hydrothermal process: High adsorption capability to azo dyes, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 408 (2012) 79-86. [27] A. Maerle, I. Kasyanov, I. Moskovskaya, B. Romanovsky, Mesoporous MgO: Synthesis, physicochemical, and catalytic properties, Russian Journal of Physical Chemistry A, 90 (2016) 1212-1216.
16
[28] C. Takai, H. Watanabe, T. Asai, M. Fuji, Determine apparent shell density for evaluation of hollow silica nanoparticle, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 404 (2012) 101105. [29] M. Mantilaka, H. Pitawala, R. Rajapakse, D. Karunaratne, K.U. Wijayantha, Formation of hollow bone-like morphology of calcium carbonate on surfactant/polymer templates, Journal of Crystal Growth, 392 (2014) 52-59. [30] G. Yan, J. Huang, J. Zhang, C. Qian, Aggregation of hollow CaCO 3 spheres by calcite nanoflakes, Materials Research Bulletin, 43 (2008) 2069-2077. [31] M.S. Mastuli, N.S. Ansari, M.A. Nawawi, A.M. Mahat, Effects of cationic surfactant in sol-gel synthesis of nano sized magnesium oxide, APCBEE Procedia, 3 (2012) 93-98. [32] Z.-S. Chao, E. Ruckenstein, Effect of the nature of the templating surfactant on the synthesis and structure of mesoporous V-Mg-O, Langmuir, 18 (2002) 734-743. [33] S. Shahkarami, A.K. Dalai, J. Soltan, Enhanced CO2 Adsorption Using MgO-Impregnated Activated Carbon: Impact of Preparation Techniques, Industrial & Engineering Chemistry Research, 55 (2016) 5955-5964. [34] M. Ahmed, Z. Abou-Gamra, Mesoporous MgO nanoparticles as a potential sorbent for removal of fast orange and bromophenol blue dyes, Nanotechnology for Environmental Engineering, 1 (2016) 10. [35] J. Hu, Z. Song, L. Chen, H. Yang, J. Li, R. Richards, Adsorption properties of MgO (111) nanoplates for the dye pollutants from wastewater, Journal of Chemical & Engineering Data, 55 (2010) 3742-3748. [36] J.F. Hernández-Sierra, F. Ruiz, D.C.C. Pena, F. Martínez-Gutiérrez, A.E. Martínez, A.d.J.P. Guillén, H. Tapia-Pérez, G.M. Castañón, The antimicrobial sensitivity of Streptococcus mutans to nanoparticles of silver, zinc oxide, and gold, Nanomedicine: Nanotechnology, Biology and Medicine, 4 (2008) 237-240. [37] J. Sawai, T. Yoshikawa, Quantitative evaluation of antifungal activity of metallic oxide powders (MgO, CaO and ZnO) by an indirect conductimetric assay, Journal of applied microbiology, 96 (2004) 803-809. [38] S. Makhluf, R. Dror, Y. Nitzan, Y. Abramovich, R. Jelinek, A. Gedanken, Microwave‐Assisted Synthesis of Nanocrystalline MgO and Its Use as a Bacteriocide, Advanced Functional Materials, 15 (2005) 1708-1715. [39] J. Bico, C. Tordeux, D. Quéré, Rough wetting, EPL (Europhysics Letters), 55 (2001) 214. [40] P. Jeevanandam, K. Klabunde, Redispersion and reactivity studies on surfactant-coated magnesium oxide nanoparticles, Langmuir, 19 (2003) 5491-5495. [41] Y. Liu, D. Hou, G. Wang, A simple wet chemical synthesis and characterization of hydroxyapatite nanorods, Materials Chemistry and Physics, 86 (2004) 69-73. [42] K. Lin, J. Chang, R. Cheng, M. Ruan, Hydrothermal microemulsion synthesis of stoichiometric single crystal hydroxyapatite nanorods with mono-dispersion and narrow-size distribution, Materials Letters, 61 (2007) 1683-1687. [43] Y. Wu, S. Bose, Nanocrystalline hydroxyapatite: micelle templated synthesis and characterization, Langmuir, 21 (2005) 3232-3234.
17
[44] P. Andreozzi, S.S. Funari, C. La Mesa, P. Mariani, M.G. Ortore, R. Sinibaldi, F. Spinozzi, Multi-to unilamellar transitions in catanionic vesicles, The Journal of Physical Chemistry B, 114 (2010) 8056-8060. [45] X. Wang, N. Zhu, B. Yin, Preparation of sludge-based activated carbon and its application in dye wastewater treatment, Journal of hazardous materials, 153 (2008) 22-27. [46] A.M. Manisha Sharma, Akansha Mehta, Diptiman Choudhury, and Soumen Basu, Effect of Surfactants on the Structure and Adsorption Efficiency of Hydroxyapatite Nanorods, Journal of Nanoscience and Nanotechnology, 17 (2017) 1-11. [47] S. Chatterjee, S. Chatterjee, B.P. Chatterjee, A.K. Guha, Adsorptive removal of congo red, a carcinogenic textile dye by chitosan hydrobeads: Binding mechanism, equilibrium and kinetics, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 299 (2007) 146-152. [48] F. Çolak, N. Atar, A. Olgun, Biosorption of acidic dyes from aqueous solution by Paenibacillus macerans: Kinetic, thermodynamic and equilibrium studies, Chemical Engineering Journal, 150 (2009) 122-130. [49] M. Ghaedi, A. Hassanzadeh, S.N. Kokhdan, Multiwalled carbon nanotubes as adsorbents for the kinetic and equilibrium study of the removal of alizarin red S and morin, Journal of Chemical & Engineering Data, 56 (2011) 2511-2520. [50] M.J. Iqbal, M.N. Ashiq, Adsorption of dyes from aqueous solutions on activated charcoal, Journal of Hazardous Materials, 139 (2007) 57-66. [51] A.A. Inyinbor, F.A. Adekola, G.A. Olatunji, Kinetics, isotherms and thermodynamic modeling of liquid phase adsorption of Rhodamine B dye onto Raphia hookerie fruit epicarp, Water Resources and Industry, 15 (2016) 14-27. [52] A.-H. Chen, C.-Y. Yang, C.-Y. Chen, C.-Y. Chen, C.-W. Chen, The chemically crosslinked metalcomplexed chitosans for comparative adsorptions of Cu (II), Zn (II), Ni (II) and Pb (II) ions in aqueous medium, Journal of Hazardous Materials, 163 (2009) 1068-1075. [53] A.S. Özcan, B. Erdem, A. Özcan, Adsorption of Acid Blue 193 from aqueous solutions onto BTMAbentonite, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 266 (2005) 73-81. [54] O. Choi, K.K. Deng, N.-J. Kim, L. Ross, R.Y. Surampalli, Z. Hu, The inhibitory effects of silver nanoparticles, silver ions, and silver chloride colloids on microbial growth, Water research, 42 (2008) 3066-3074.
18
Table1: Detailed composition of the materials used for the synthesis of mesoporous MgO Sample Name
Concentration MgCl2 (M)
M-0.125
1
M-0.25
1
M-0.5
of Concentration C16TAB (M)
of
0.125
Concentration of Concentration SDS (M) of NaOH (M) 0.125
2
0.25
0.25
2
1
0.50
0.5
2
M-1
1
1.00
1.0
2
M-1.5
1
1.50
1.5
2
Table 2: Nitrogen adsorption-desorption parameters for mesoporous MgO showing the effect of different molar concentration of catanionic surfactant. Sample Name
Ratio of SDS(M)
C16TAB: Surface Area ( Pore Volume 2 -1 3 -1 mg ) (cm g )
Pore (nm)
M-0.125
0.125:0.125
165
1.45
29.3
M-0.25
0.25:0.25
141
1.28
27.2
M-0.50
0.5:0.5
138
1.19
19.6
M-1.0
1.0:1.0
123
0.70
15.4
M-1.5
1.5:1.5
109
0.50
12.2
Diameter
Table 3: Nitrogen adsorption-desorption parameters for mesoporous MgO nanorods by varying the chain length of a cationic surfactant. Sample Name
Ratio CnTAB:SDS
C12TAB : SDS
0.125:0.125
C14TAB:SDS
of Surface Area m2g-1 )
(
Pore Volume (cm3g-1)
Pore (nm)
50
0.73
40
0.125:0.125
110
0.97
32
C16TAB:SDS
0.125:0.125
165
1.45
29
C18TAB :SDS
0.125:0.125
180
1.48
19
Diameter
19
Table 4: Thermodynamic parameters at different temperatures for removal of MB, AZ and RD on MgO. Dyes
T(K)
∆G (kJ mol−1)
∆H (kJ mol−1)
∆S (J mol K−1)
MB
293
-4.80
-6.71
6.47
313
-4.67
333
-4.54
353
-4.41
293
-6.84
-8.84
6.80
313
-6.71
333
-6.58
353
-6.43
293
-8.65
-10.91
7.70
313
-8.49
333
-8.34
353
-8.19
RD
AZ
Table 5: Kinetic Parameter for adsorption of dyes on MgO Dyes
Pseudo-first-order
Pseudo-second-order
Intra-Particle diffusion
K1 @
qe #
R2
K2 +
qe #
R2
Kdiff
C
R2
MB
0.075
388
0.91
1×10-5
400
0.96
4.18
25.54
0.87
AZ
0.019
57.6
0.89
1.8×10-5
132
0.98
3.54
16.73
0.90
RD
0.021
140
0.92
6×10-5
285
0.97
5.25
5.95
0.87
20
Table 6: Langmuir and Freundlich constants and correlation coefficients for adsorption of MB, AZ, RD on MgO. Dyes
Freundlich
Langmuir
Kf *
N
R2
2
Q 0#
B
R2
2
Blank MgO
20.0
3.00
0.40
10.44
30.0
0.05
0.70
4.9
MB
160.78
1.99
0.96
7.07
333.33
0.8
0.99
3.8
AZ
120.78
2.40
0.95
8.09
250
0.37
0.97
1.59
RD
97.72
1.41
0.93
6.47
200
0.17
0.95
2.01
21
0.20
a
Blank MgO-0.125 MgO-0.25 MgO-0.5 MgO-1 MgO-1.5
dV/dr (cm3/g/nm)
Volume adsorbed (cm3/g)
2000 1500 1000 500 0
0.0
0.2
0.4
0.6
0.8
Relative pressure (P/Po)
1.0
b
Blank MgO-0.125 MgO-0.25 MgO-0.50 MgO-1 MgO-1.5
0.15
0.10
0.05
0.00
0
20
40
60
80
100
Pore diameter (nm)
Fig 1: (a) Nitrogen sorption plots and (b) Pore size distribution of synthesized blank MgO and by different ratios of surfactants, C16TAB and SDS (0.125, 0.25, 0.5, 1, 1.5 M)
22
a
b
100 nm
100nm
c
d
100 nm
100 nm
e
Fig 2: (a-b) FESEM images, (c-d) HRTEM images and (e) EDS spectrum of MgO with different surfactants ratio (0.125M for (a)/(c) and 1.5 M for (b)/(d)).
23
+
NaOH
MgCl2
Cat-anionic surfactant
MgO (Rod shape)
Fig 3. Mechanism showing the synthesis of mesoporous MgO nanorods.
C12TAB C16TAB
a
C14TAB C18TAB
600
400
200
0 0.0
0.2
0.4
0.6
0.8
C18TAB C14TAB
0.15
dV/dr (cm3/g/nm)
Volume adsorbed (cm3/g)
800
1.0
C16TAB C12TAB
b
0.10
0.05
0.00
Relative pressure (P/Po)
10
20
30
40
50
60
70
80
90 100
Pore diameter (nm)
Fig 4: (a) Nitrogen sorption plots and (b) Pore size distribution plots for MgO-0.125 with the different chain length of CnTAB (where n= 12-18).
24
bB 300nm
aA
200 nm 200 nm
200 200nm nm
c
d
100 nm
100 nm
e
Fig. 5: (a-b) FESEM images, (c-d) HRTEM images and (e) EDS spectrum of MgO-0.125 with different chain length of CnTAB ((a)/(c) for C12TAB and (b)/(d) for C18TAB).
25
Intensity (a.u)
6000 5000
110
a
(200)
(220)
4000 3000
(111)
2000 1000 0
Blank MgO MgO-0.125(C18TAB)
105
% Transmittance
Blank C12TAB C14TAB C16TAB C18TAB
7000
b
100 95 90 85
3345
1431.7
80 75 70
20
30
40
50
60
70
578 MgO
1095
4000
3500
2 (degree)
3000
2500
2000
Wavenumber
1500
1000
500
(cm-1)
Fig.6 (a) X-ray diffraction pattern of MgO synthesized by different chain length of C nTAB (where n= 12-18) and (b) FTIR spectra of blank MgO and MgO-0.125 (C18TAB).
110
90
100
MB AZ RD Blank MgO
80
70 60 50
60 40 20
40 30 2
3
4
5
6
7
8
9
10
0
11
0
20
pH 100 90 80
c
MB AZ RD Blank MgO
70 60 50 40 30 20 10 100
150
200
250
40
60
80
100
120
Contact time (min)
ln(qe/Ce)
% Adsorption
b
80
qe(mg/g)
% Adsorption
a
MB AZ RD Blank
100
300
350
Dye Concentration (mg/L)
400
5.0 4.9 4.8 4.7 4.6 4.5 4.4 4.3 4.2 4.1 4.0 3.9 3.8
d
MB AZ RD
0.28
0.29
0.30
0.31
0.32
0.33
0.34
0.35
1/T
Fig 7: Plots for (a) effect of pH, (b) effect of time, (c) effect of concentration and (d) Van’t Hoff plots for the adsorption of MB, AZ, RD on MgO-0.125 (C18TAB).
26
7.0
2
AZ, R = 0.97
6.0
1.0
2
RD, R = 0.98
5.5
b
MB, R2 = 0.91 AZ, R2 = 0.98
2
2
RD, R = 0.97
5.0
t/qt
ln (qt-qe) (mg/g)
a
MB, R = 0.99
6.5
4.5
0.5
4.0 3.5 3.0 2.5
0.0
10
20
30
40
50
60
70
80
10
90 100
20
30
40
100
MB AZ RD
90 80
50
60
70
80
90 100
Time (min)
Time (min)
c
qt
70 60 50 40 30 20
‘
3
4
5
6
7
8
9
10
11
12
1/2
t
Fig 8. Plots for (a) pseudo- first order, (b) pseudo-second order, (c) Weber Moris plot explaining mechanism of adsorption of MB, AZ, RD on MgO.
27
\\ 1.2
MB, y = 0.0057x + 0.1003,R² = 0.98
1.1 0.9
log Ce
0.6 0.5 0.4
2.2 2.0 1.8
0.3
1.6
0.2
1.4
0.1 10
20
30
40
Ce
50
60
RD, y = 0.8778x + 2.1293,R² = 0.9851
2.4
0.7
70
b
AZ, y = 0.7355x + 1.8881,R² = 0.9806
2.6
RD, y = 0.0087x + 0.171,R² = 0.97
0.8
MB,y = 0.5454x + 2.1812,R² = 0.9713
2.8
AZ, y = 0.0046x + 0.0639,R² = 0.99
1.0
Ce/Qe
3.0
a
1.2 -1.0
-0.8
-0.6
-0.4
-0.2
0.0
log Ce
Fig 9: (a) Langmuir isotherm and (b) Freundlich isotherm for the adsorption of MB, AZ, RD on MgO-0.125 (C18TAB).
100
MB AZ RD
% Adsoprtion
80 60 40 20 0
1 2 3 4 Fig 10 : Plot showing percentage removal vs. no. of cycles
5
No. of Cycles
28
% Growth of Bacteria
100 E. coli B. subtilis 80
60
40
20
0 0
100
200
300
400
500
600
Concentration (g)
Fig 11. Concentration vs percentage bacterial growth plot for antimicrobial activity of MgO0.125 (C18TAB) using E. coli and B. subtilis.
29