Kaouthar
Ahmouda
*ab,
Moussa
Boudiaf
ac and
Boubaker
Benhaoua
d
aDepartment of Process Engineering and Petrochemistry, Faculty of Technology, University of El Oued, El Oued, 39000, Algeria. E-mail: ahmouda-kaouthar@univ-eloued.dz; drssmoussa@yahoo.fr
bRenewable Energy Research Unit in Arid Zones, University of El Oued, El Oued, 39000, Algeria
cLCIMN, Laboratory, Department of Process Engineering, Faculty of Technology, University Ferhat, Abbas Setif, 19000, Sétif, Algeria. E-mail: benhaouab@yahoo.fr
dDepartment of Physics, Faculty of Exact Sciences, University of El Oued, El Oued, 39000, Algeria
First published on 13th July 2022
In this paper, the adsorption of Evans blue (EB) and methyl orange (MO) azo dyes on four greenly synthesized magnetite nanoparticles has been studied to investigate the effect of the mediating plant extract's acidity on magnetite surface reactivity in azo dye adsorption. Magnetite surface reactivity has been studied through the analysis of preferential attachment of dye chromophore and auxochrome groups on magnetite nanoparticles, and adsorption yields. According to the contents of chromophore and auxochrome groups in dye structures, the mediating plant extract's acidity effect on acid site types and densities was also deduced. Used plants for the green synthesis were: Artemisia herba-alba (L), Matricaria pubescens (L), Juniperus phoenicea (L), and Rosmarinus officinalis (L), and their extract pHs were respectively 5.25, 5.05, 4.63, and 3.69. The four greenly synthesized samples of magnetite were characterized by XRD, SEM, ATR-FTIR, and UV-Vis techniques. The novelty of this paper lies in highlighting the influence of the mediating plant extract's acidity on the greenly synthesized magnetite surface reactivity towards the preferential attachment of chromophore and auxochrome functional groups in azo dye adsorption, where obtained results show that the mediating plant extract's acidity has a clear effect on the preferential attachment of chromophore and auxochrome groups on magnetite surfaces as well as on azo dyes' adsorption yields and capacities. Indeed, the decrease in the plant extract's acidity leads to an increase in the attachment of chromophore groups and a decrease in the attachment of auxochrome groups. So, it leads to an increase in Lewis acid site density and a decrease in Brønsted acid site density of magnetite surfaces. Also, the decrease of the plant extract's acidity leads to an increase in the studied dye adsorption yields, and this is because the majority of functional groups of MO and EB dyes are chromophores that attach to Lewis acid sites. The difference found in adsorption yields of EB and MO on all four magnetite samples is due to the fact that the ratio of chromophore/auxochrome groups in EB is remarkably greater than that in MO. The linear and non-linear pseudo-first-order and pseudo-second-order kinetics of the adsorption as well as the intra-particle diffusion mechanism have been analyzed. Obtained results indicate that in all adsorption processes the adsorption kinetics followed a linear pseudo-first-order kinetic model, and film diffusion is the step that controlled adsorption mechanisms. The thermodynamic studies of EB and MO adsorption processes on the four magnetite surfaces have been analyzed in the temperature range of 303.15–318.15 K. Obtained results reveal the endothermic nature of the adsorption in all cases.
Dyes are composed of chromophores that are commonly electron withdrawing and auxochromes that are usually electron-releasing groups.4 The most important chromophores, as defined in this way, are: NN, CO, –CHN, NO2, NO, NOH, CN, CN, CC, and CC groups and ionizing auxochromes mainly include: SO3H, OH, COOH, NH2, NH3, NHCH3, and N(CH3)2 groups.4,5
Adsorption is one of the most effective processes for the removal and recovery of colored materials and dyes from effluents.6–9 Nanomaterials are widely used in the purification of aqueous media due to their advantages, such as high surface area and increased number of active sites.10–13 They therefore allow a rapid thermodynamic equilibrium between adsorbent and adsorbate during the adsorption process and selective removal of pollutants.7–9 Several factors influence the adsorption process, mainly the solution chemistry,7,14 the characteristics of the dye (adsorbate)7,15 and the adsorbent surface properties.16–22 Saha et al.7 studied the preferential adsorption of seven different dyes on magnetite NPs. They reported that the magnetite surface preferred adsorbing dyes containing higher OH content. Xiao et al.15 studied the preferential adsorption of different cationic and anionic dyes on iron nanoparticles. They reported that iron NPs preferred removing cationic dyes more than anionic dyes.
Other authors studied the effect of changing the adsorbent surface properties by binding ligands on the adsorbent surface. Khurshid et al.17 found that the use of amine-functionalized cobalt–iron NP surface enhanced the removal of anionic azo dyes. Mahmoodi et al.22 synthesized a titania/silica nano-hybrid (TSNH) and an amine-functionalized titania/silica nano-hybrid (AFTSNH) to use them in Reactive Red 198 and Acid Red 14 removal from wastewater. They found that the AFTSNH adsorbent showed high dye adsorption capacities compared to the TSNH adsorbent. The authors of ref. 23 prepared silica nanoparticles (SN) and amine-functionalized silica nanoparticles (AFSN) and then used them in Acid Red 14, Acid Black 1 and Acid Blue 25 removal. They reported that AFSN preferred adsorbing the studied dyes than SN. Meanwhile, the study of Madrakian et al.24 reported that magnetite-modified activated carbon preferred adsorbing cationic dyes than anionic dyes. Moreover, a comparative study on adsorption of methylene blue on sericin-modified and unmodified magnetite NPs25 reported that sericin-modified magnetite NPs were approximately 40% more effective than the unmodified magnetite NPs. Furthermore, in the study of the preferential adsorption of magnetite NP loaded fig leaves (MNLFL) and magnetite NP loaded azolla (MNLA) to remove crystal violet and methylene blue,26 the authors found that MNLFL preferred adsorbing crystal violet more than MNLA.
Surface acid site types and densities were also found to impact the adsorption process. Indeed, a study of the adsorption of organic contaminants on both natural and synthesized magnetite20 found that the adsorption on natural magnetite was more efficient than that on the synthesized one, and this is because of its higher surface site density. Moreover, Gogoi et al.27 studied the degradation of catechol using an Fe3O4–CeO2 nanocomposite as a Fenton-like heterogeneous catalyst. They reported that the increase of Brønsted acid site density of this nanocomposite increased the degradation of catechol.
The impact of changing the mediating plants on the behavior of greenly synthesized metal oxide NPs in dye adsorption was studied in several works. The effect of three different tea extracts on the capacity of greenly synthesized iron oxide surfaces in the removal of methyl green dye from aqueous solution has been studied by Huang et al.16 They reported that the plant extract had an effect on the adsorption yields of methyl green dye on the three iron oxide NPs, where yields varied from 81.2% to 75.6 to 67.1%. Duyen et al.28 synthesized metal oxide NPs using the extracts of flowers, bark, and leaf of Tecoma stans in order to use them in the removal of Congo red (CR) and crystal violet (CV) dyes. They reported that the adsorbent derived from flower extract gave better adsorption efficiency than those derived from other extracts. Islam et al.29 synthesized magnetite NPs using six plant extracts in order to use them in the removal of methyl orange (MO) and crystal violet (CV) dyes. They reported that plant extract had an effect on magnetite NP surface reactivity in the adsorption, where magnetite NPs synthesized using tea extract showed the highest performance (MO 92.34%, CV 96.1%). Ahmouda et al.30 used different greenly synthesized magnetite NPs in the removal of methyl green (MG) dye via the adsorption process. They reported that the mediating plant extract's acidity had an effect on the preferential adsorption of MG on the magnetite NP surface. Indeed, magnetite NPs synthesized using plant extract having the lowest acidity exhibited the highest acid site density (OH groups) and hence the highest performance in the removal of MG.
As it is known that the adsorption of dyes from wastewater is directly affected by the reactivity of the adsorbent surface towards the attachment of dyes' functional groups, looking for parameters that could control the greenly synthesized magnetite NP surface reactivity is always of importance. In this way, this paper looks for mediating plant extract parameters that could, in the case of green synthesis, impact the magnetite NP surface reactivity in azo dye adsorption.
Azo dyes are the largest and most versatile class of organic dyes.31 These dyes contain one or more azo bonds (NN). The complex aromatic structures of azo dyes make them more stable and more difficult to remove from the effluents discharged into water bodies.31 The used dyes in this study are: methyl orange (MO) and Evans blue (EB) azo dyes, and their chemical structures are illustrated in Fig. 1. The EB molecule is composed of chromophore groups such as benzene, phenyl, phenyldiazonium, toluene, NN, CC, and CN groups linked to the benzene ring, and of auxochrome groups such as sulphonic acid (SO−3), phenol, and aniline groups. Meanwhile, the MO molecule is composed of chromophore groups such as benzene, phenyl, phenyldiazonium, NN, CC, and CN, and of auxochrome groups such as sulphonic acid (SO−3) and dimethylamine (N(CH3)2).
The attachment of these functional groups on the adsorbent surface is based on functional group properties (chromophore or auxochrome) towards the surface acid site type (Lewis or Brønsted). In the case of EB and MO dyes, SO−3,32 dimethylamine,33 phenol,34 and aniline35 auxochrome groups are electron donating, and their attachment on the adsorbent surface is based on their ionic interaction with the developed positively charged Brønsted acid sites of the surface. Meanwhile, the chromophore groups phenyldiazonium, phenyl,36 benzene, NN, CC, CN,32 and toluene37 are electron withdrawing, and their attachment on the adsorbent surface is based on their interactions with the Lewis acid sites of the adsorbent surface.
In this paper, the effect of the mediating plant extract's acidity on the greenly synthesized magnetite surface reactivity in the adsorption of methyl orange (MO) and Evans blue (EB) azo dyes has been investigated through the analysis of MO and EB adsorption on four greenly synthesized magnetite samples, with the aim of studying the effect of the mediating plant extract's acidity on the preferential attachment of the studied dye chromophores and auxochromes (functional groups) on greenly synthesized magnetite NPs, and adsorption yields. For this purpose, after the accomplishment of adsorption experiments, the free functional groups that are not attached on magnetite surfaces have been deeply analyzed in all dye residual solutions using ATR-FTIR spectroscopy, so as to perceive their preferential attachment on the four magnetite surfaces. Based on the analysis of preferential attachments of chromophore and auxochrome groups, it was possible to compare between Brønsted and Lewis acid site densities on each magnetite surface and their influence on adsorption yields and capacities. To the best of our knowledge, there is no study in the literature that has dealt with the influence of the mediating plant extract's acidity on magnetite surface reactivity towards the preferential attachment of dye functional groups, and thus on surface acid site type and density. The used plants are Artemisia herba-alba (L), Matricaria pubescens (L), Juniperus phoenicea (L), and Rosmarinus officinalis (L). Their extract pHs are respectively 5.25, 5.05, 4.63, and 3.69. The synthesized Fe3O4 samples are respectively denoted in this paper by ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4 and JUN–Fe3O4. They were characterized by XRD, SEM, FTIR-ATR, and UV-Vis techniques.
All experiment sets are sonicated in an ultrasonic bath for 15 minutes and they were then stirred continually for 60 minutes until a steady state was reached. All adsorption experiments were carried out under ambient conditions in batch mode, and they were performed in triplicate for data consistency.
Kinetic experiments were performed by withdrawing samples of dye/Fe3O4 solutions at a regular time interval to obtain, after centrifugation, adequate aliquots for the purpose of quantifying residual dye concentrations and the adsorbed amounts. The concentrations of residual dye aqueous solutions were quantified using a UV-Vis spectrophotometer at absorbance maxima of EB (λmax = 602 nm) and MO (λmax = 463 nm). Furthermore, the adsorbed amounts of EB and MO molecules are calculated from the calibration curve for all adsorption experiments (Y = 67.02X + 0.0442, R2 = 0.9987 and Y = 31.39X + 0.0346, R2 = 0.9985, respectively). On another side, after the adsorption was accomplished and steady state reached, the aliquots were centrifuged to separate liquid solutions and solid phases. The liquid solutions, which represent MO/magnetite and EB/magnetite residual solutions, were then analyzed using ATR-FTIR spectroscopy.
To calculate the adsorption capacity (qe in mg g−1) and the amount of MO and EB ions adsorbed per unit mass (qt in mg g−1) of JUN–Fe3O4, MAT–Fe3O4, ROS–Fe3O4 and ARM–Fe3O4 at equilibrium contact time, the following equations were used:
(1) |
(2) |
Adsorption yield was calculated using the following equation:
(3) |
ln(qe − qt) = lnqe − K1t | (4) |
qt = qe(1 − e−K1t) | (5) |
If the active surface of the adsorbent is regarded as invariable, the reaction could be treated as pseudo-first-order kinetic. However, once the active sites have been saturated, the transfer at the adsorbate/adsorbent particle interface may be limited by mass transfer.44
The pseudo second-order (PSO) model is proposed by Ho and McKay.40 It is based on the adsorption capacity expressed as follows:
(6) |
(7) |
(8) |
(9) |
A further formula manipulation gives the following:
(10) |
Additionally, eqn (11) can be used when the rate of adsorption is controlled by liquid film diffusion.46
(11) |
−ln(1 − F) = kft | (12) |
The values of kf = 3DfCe/r0δCr for the adsorption of MO and EB on ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4, and JUN–Fe3O4 are obtained from the slopes of the fitted lines (plots of −ln(1 − F) vs. time), and the values of effective diffusion coefficient, Df (m2 s−1), can then be obtained from Df = kfr0Cr/3Ce.
The linearity test of Boyd plots −ln(1 − F) and −ln(1 − F2) versus time plots is employed to distinguish between the film diffusion and particle diffusion-controlled adsorption mechanism. If the plot of −ln(1 − F) versus time is a straight line passing through the origin, then the adsorption rate is governed by the particle diffusion mechanism, otherwise if −ln(1 − F2) versus time is a straight line passing through the origin then the adsorption is governed by film diffusion.
To calculate the desorption yields (R%) of MO and EB ions at contact time t = 60 min from JUN–Fe3O4, MAT–Fe3O4, ROS–Fe3O4 and ARM–Fe3O4 surfaces, the following equation was used:
(13) |
The adsorption capacity, qeT (mg g−1), was calculated using the following equations:
(14) |
Adsorption yield was calculated using the following equation:
(15) |
The activation enthalpy (ΔH0) of EB and MO adsorption processes on the magnetite NP surface was determined using the Arrhenius equation as follows:
(16) |
(17) |
The values of activation enthalpy ΔH0 (kcal mol−1) and entropy ΔS0 (cal mol−1 K−1) were respectively calculated from the slope and intercept of plots between lnkd and 1/T. ΔG0 (kcal mol−1) was then calculated using the following equation:
ΔG0 = −RTlnkd | (18) |
The free energy change indicates the degree of spontaneity of the adsorption process. The higher negative value reflects more energetically favorable adsorption. The activation energy, ΔEa (kcal mol−1), of EB and MO adsorption processes on magnetite surfaces was determined using Arrhenius's equation:
(19) |
(20) |
Fig. 2 XRD patterns of (A) ROS–Fe3O4, (B) ARM–Fe3O4, (C) MAT–Fe3O4, and (D) JUN–Fe3O4 NPs (JCPDS file 01-076-0958). |
The X-ray diffraction pattern in Fig. 2(B) exhibits Bragg reflection peaks at around 2θ = 16.20°, 16.70°, 20.39°, 22.42°, 29.75°, 30.80°, 32.30°, 41.10°, 42.53°, 49.82°, and 52.72°. All Bragg peaks are in agreement with orthorhombic Fe3O4 powder and correspond to Miller indices 021, 210, 212, 030, 400, 041, 106, 251, 522, 534, and 644, respectively (JCPDS file 01-076-0958).
The X-ray diffraction pattern in Fig. 2(C) exhibits Bragg reflection peaks at around 2θ = 16.20°, 22.56°, 26.04°, 32.28°, 37.11°, 41.59°, 49.98°, and 52.69°. All Bragg peaks are in agreement with orthorhombic Fe3O4 powder and correspond to Miller indices 021, 030, 400, 106, 404, 251, 534, and 644, respectively (JCPDS file 01-076-0958).
The X-ray diffraction pattern in Fig. 2(D) exhibits Bragg reflection peaks at around 2θ = 16.35°, 20.58°, 22.60°, 25.77°, 29.94°, 32.47°, 41.59°, 42.69°, 49.98°, and 52.69°. All Bragg peaks are in agreement with orthorhombic Fe3O4 powder and correspond to Miller indices 021, 212, 030, 400, 001, 106, 251, 522, 534, and 644, respectively (JCPDS file 01-076-0958).
The average diameters of different magnetite samples, presented in Table 1, are calculated from XRD patterns using Scherrer's equation:48
(21) |
Samples | Average diameter (nm) |
---|---|
ARM–Fe3O4 | 41.49 |
ROS–Fe3O4 | 39.89 |
MAT–Fe3O4 | 33.13 |
JUN–Fe3O4 | 29.27 |
Fig. 3 FTIR spectra of the synthesized (A) ARM–Fe3O4, (B) ROS–Fe3O4, (C) MAT–Fe3O4, and (D) JUN–Fe3O4 samples. |
Sample | O–H cm−1 | C–H cm−1 | CC cm−1 | C–O–C cm−1 | Fe–O cm−1 |
---|---|---|---|---|---|
ARM–Fe3O4 | 3266.69 | 2932.06 | 1590.07 | 1036.36 | 592.64 |
ROS–Fe3O4 | 3249.77 | 2930.18 | 1590.83 | 1038.75 | 591.83 |
MAT–Fe3O4 | 3235.57 | 2929.75 | 1591.21 | 1039.54 | 592.46 |
JUN–Fe3O4 | 3223.41 | 2928.82 | 1594.63 | 1039.45 | 592.69 |
Fig. 3 shows that the peaks of hydroxyl groups appear in remarkably different areas. Meanwhile, the hydroxyl group peak area appears to be the broadest on the ARM–Fe3O4 surface, next on ROS–Fe3O4, then on MAT–Fe3O4, and finally on JUN–Fe3O4. This reveals that the density of functional OH groups is higher on the ARM–Fe3O4 surface, next on ROS–Fe3O4, then on MAT–Fe3O4, and finally on JUN–Fe3O4.
αhν = A(hν − Eg)n | (22) |
E g of the direct transition of all samples were obtained from plotting (αhν)2 as a function of αhν by the extrapolation of the linear portion of the curve (Fig. 4). However, Eg of the indirect transition of all samples were obtained from plotting (αhν)1/2 as a function of αhν by the extrapolation of the linear portion of the curve (Fig. 5).
Fig. 4 Plots of (αhν)2versus (αhν) for direct transition of synthesized Fe3O4 samples sonicated in acetone for 15 minutes. |
Fig. 5 Plots of (αhν)1/2versus (αhν) for indirect transition of synthesized Fe3O4 samples sonicated in acetone for 15 minutes. |
Estimated direct band gap energies of JUN–Fe3O4, MAT–Fe3O4, ROS–Fe3O4, and ARM–Fe3O4 samples were found to be 2.97, 2.96, 2.95 and 2.87 eV, respectively, which are close to that found by El Ghandoor et al.52 They found that direct gap energy for Fe3O4 equals Eg = 2.87 eV.
The estimated indirect band gap energies of ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4 and JUN–Fe3O4 phases were found to be 2.51, 2.55, 2.60 and 2.64 eV, respectively, which are higher than reported by other authors.52 They found that indirect gap energy for Fe3O4 equals Eg = 1.92 eV.
It is clear that the direct gap energy is closer to the theoretical value than the indirect gap energy. The values of all direct band gap energies of magnetite samples classify them as semiconductors. The energy band gap of the semiconductors is between 0 and 3 eV.53
Fig. 6 SEM images of greenly synthesized (a) JUN–Fe3O4, (b) MAT–Fe3O4, (c) ROS–Fe3O4 and (d) ARM–Fe3O4 NPs. |
Fig. 7 Adsorption capacities of (a) EB and (b) MO on ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4, and JUN–Fe3O4 surfaces. Error bars represent the standard deviation of three replicates. |
Sample | t (min) | q t (mg g−1) | STD | t (min) | q t (mg g−1) | STD | t (min) | q t (mg g−1) | STD | t (min) | q t (mg g−1) | STD |
---|---|---|---|---|---|---|---|---|---|---|---|---|
EB/ARM–Fe3O4 | 05 | 09.15 | 0.73 | 10 | 13.54 | 0.64 | 15 | 17.29 | 0.71 | 20 | 20.55 | 0.59 |
EB/ROS–Fe3O4 | 05 | 08.03 | 0.64 | 10 | 11.92 | 0.59 | 15 | 15.06 | 0.69 | 20 | 17.69 | 0.72 |
EB/MAT–Fe3O4 | 05 | 04.62 | 0.75 | 10 | 08.24 | 0.72 | 15 | 11.03 | 0.73 | 20 | 13.90 | 0.78 |
EB/JUN–Fe3O4 | 05 | 04.05 | 0.74 | 10 | 06.86 | 0.59 | 15 | 09.40 | 0.62 | 20 | 11.83 | 0.78 |
EB/ARM–Fe3O4 | 25 | 22.82 | 0.81 | 30 | 25.22 | 0.59 | 35 | 25.29 | 0.65 | 40 | 25.33 | 0.52 |
EB/ROS–Fe3O4 | 25 | 20.84 | 0.63 | 30 | 22.75 | 0.57 | 35 | 22.80 | 0.52 | 40 | 22.84 | 0.72 |
EB/MAT–Fe3O4 | 25 | 16.25 | 0.64 | 30 | 17.87 | 0.78 | 35 | 17.95 | 0.67 | 40 | 17.99 | 0.56 |
EB/JUN–Fe3O4 | 25 | 13.87 | 0.68 | 30 | 16.71 | 0.65 | 35 | 16.76 | 0.72 | 40 | 16.78 | 0.71 |
EB/ARM–Fe3O4 | 45 | 25.35 | 0.66 | 50 | 25.37 | 0.74 | 60 | 25.39 | 0.78 | |||
EB/ROS–Fe3O4 | 45 | 22.87 | 0.52 | 50 | 22.90 | 0.53 | 60 | 22.92 | 0.71 | |||
EB/MAT–Fe3O4 | 45 | 18.00 | 0.81 | 50 | 18.01 | 0.61 | 60 | 18.03 | 0.63 | |||
EB/JUN–Fe3O4 | 45 | 16.79 | 0.80 | 50 | 16.81 | 0.71 | 60 | 16.83 | 0.74 |
Sample | t (min) | q t (mg g−1) | STD | t (min) | q t (mg g−1) | STD | t (min) | q t (mg g−1) | STD | t (min) | q t (mg g−1) | STD |
---|---|---|---|---|---|---|---|---|---|---|---|---|
MO/ARM–Fe3O4 | 05 | 04.78 | 0.65 | 10 | 08.44 | 0.65 | 15 | 12.88 | 0.55 | 20 | 15.93 | 0.66 |
MO/ROS–Fe3O4 | 05 | 04.42 | 0.89 | 10 | 06.99 | 0.55 | 15 | 11.39 | 0.59 | 20 | 14.45 | 0.50 |
MO/MAT–Fe3O4 | 05 | 03.80 | 0.61 | 10 | 05.67 | 0.63 | 15 | 09.07 | 0.54 | 20 | 12.06 | 0.57 |
MO/JUN–Fe3O4 | 05 | 02.90 | 0.59 | 10 | 04.57 | 0.52 | 15 | 06.84 | 0.52 | 20 | 08.87 | 0.68 |
MO/ARM–Fe3O4 | 25 | 19.70 | 0.69 | 30 | 20.70 | 0.54 | 35 | 20.73 | 0.70 | 40 | 20.75 | 0.69 |
MO/ROS–Fe3O4 | 25 | 18.32 | 0.52 | 30 | 19.70 | 0.61 | 35 | 19.73 | 0.58 | 40 | 19.75 | 0.54 |
MO/MAT–Fe3O4 | 25 | 13.54 | 0.69 | 30 | 14.70 | 0.55 | 35 | 14.71 | 0.53 | 40 | 14.73 | 0.58 |
MO/JUN–Fe3O4 | 25 | 10.89 | 0.52 | 30 | 11.66 | 0.51 | 35 | 11.68 | 0.78 | 40 | 11.70 | 0.52 |
MO/ARM–Fe3O4 | 45 | 20.79 | 0.83 | 50 | 20.81 | 0.74 | 60 | 20.82 | 0.53 | |||
MO/ROS–Fe3O4 | 45 | 19.78 | 0.85 | 50 | 19.79 | 0.74 | 60 | 19.80 | 0.50 | |||
MO/MAT–Fe3O4 | 45 | 14.74 | 0.64 | 50 | 14.76 | 0.54 | 60 | 14.78 | 0.46 | |||
MO/JUN–Fe3O4 | 45 | 11.71 | 0.58 | 50 | 11.72 | 0.56 | 60 | 11.73 | 0.49 |
Sample | q e,exp (mg g−1) | q e,cal (mg g−1) | K 1 (min−1) | R 2 | q e,cal (mg g−1) | K 2 (g mg−1 min−1) | R 2 |
---|---|---|---|---|---|---|---|
EB/ARM–Fe3O4 | 25.39 | 23.85 | 0.0008 | 0.983 | 33.84 | 0.0016 | 0.972 |
EB/ROS–Fe3O4 | 22.92 | 22.93 | 0.0011 | 0.989 | 31.52 | 0.0020 | 0.985 |
EB/MAT–Fe3O4 | 18.03 | 18.61 | 0.0009 | 0.982 | 30.94 | 0.0010 | 0.946 |
EB/JUN–Fe3O4 | 16.83 | 15.50 | 0.0007 | 0.991 | 26.11 | 0.0014 | 0.983 |
MO/ARM–Fe3O4 | 20.82 | 21.52 | 0.0008 | 0.988 | 32.25 | 0.0008 | 0.965 |
MO/ROS–Fe3O4 | 19.80 | 22.85 | 0.0009 | 0.987 | 30.76 | 0.0007 | 0.961 |
MO/MAT–Fe3O4 | 14.78 | 14.50 | 0.0007 | 0.996 | 22.50 | 0.0020 | 0.994 |
MO/JUN–Fe3O4 | 11.73 | 12.27 | 0.0008 | 0.989 | 18.47 | 0.0019 | 0.969 |
Sample | q e,exp (mg g−1) | q e,cal (mg g−1) | K 1 (min−1) | R 2 | q e,cal (mg g−1) | K 2 (g mg−1 min−1) | R 2 |
---|---|---|---|---|---|---|---|
EB/ARM–Fe3O4 | 25.39 | 27.53 | 0.077 | 0.973 | 32.56 | 0.0026 | 0.964 |
EB/ROS–Fe3O4 | 22.92 | 23.96 | 0.075 | 0.975 | 29.97 | 0.0025 | 0.955 |
EB/MAT–Fe3O4 | 18.03 | 19.31 | 0.063 | 0.975 | 25.14 | 0.0023 | 0.952 |
EB/JUN–Fe3O4 | 16.83 | 18.62 | 0.052 | 0.967 | 25.07 | 0.0018 | 0.952 |
MO/ARM–Fe3O4 | 20.82 | 22.62 | 0.060 | 0.948 | 29.88 | 0.0018 | 0.924 |
MO/ROS–Fe3O4 | 19.80 | 23.89 | 0.055 | 0.925 | 29.69 | 0.0015 | 0.923 |
MO/MAT–Fe3O4 | 14.78 | 15.79 | 0.064 | 0.952 | 21.05 | 0.0026 | 0.926 |
MO/JUN–Fe3O4 | 11.73 | 12.88 | 0.060 | 0.948 | 17.02 | 0.0030 | 0.931 |
Sample | r 0 × 10−9 (m) | k p × 10−3 (1/s) | R 2 | D p × 10−19 (m2 s−1) | k f × 10−3 (1/s) | R 2 | D f × 10−12 (m2 s−1) |
---|---|---|---|---|---|---|---|
EB/ARM–Fe3O4 | 41.94 | 2.18 | 0.963 | 3.89 | 1.44 | 0.953 | 03.38 |
EB/ROS–Fe3O4 | 39.89 | 2.26 | 0.985 | 3.68 | 1.51 | 0.962 | 05.91 |
EB/MAT–Fe3O4 | 33.13 | 2.51 | 0.977 | 2.80 | 2.08 | 0.956 | 14.97 |
EB/JUN–Fe3O4 | 29.27 | 1.89 | 0.993 | 1.65 | 1.08 | 0.998 | 08.13 |
MO/ARM–Fe3O4 | 41.94 | 1.79 | 0.990 | 3.67 | 1.30 | 0.965 | 07.67 |
MO/ROS–Fe3O4 | 39.89 | 2.12 | 0.963 | 3.43 | 1.60 | 0.930 | 10.62 |
MO/MAT–Fe3O4 | 33.13 | 1.77 | 0.977 | 1.98 | 1.32 | 0.951 | 14.28 |
MO/JUN–Fe3O4 | 29.27 | 1.75 | 0.970 | 1.46 | 1.28 | 0.970 | 18.97 |
Sample | T (K) | lnKD | lnK2 | E a (kcal mol−1) | ΔH0 (kcal mol−1) | ΔS0 (cal mol−1 K−1) | ΔG0 (kcal mol−1) |
---|---|---|---|---|---|---|---|
EB/ARM–Fe3O4 | 303.15 | 1.82 | 9.71 | 2.79 | 2.85 | 13.03 | −1.09 |
308.15 | 1.92 | 9.81 | −1.18 | ||||
313.15 | 1.98 | 9.84 | −1.23 | ||||
318.15 | 2.05 | 9.94 | −1.29 | ||||
EB/ROS–Fe3O4 | 303.15 | 1.54 | 9.43 | 3.21 | 3.32 | 14.04 | −0.93 |
308.15 | 1.65 | 9.54 | −1.00 | ||||
313.15 | 1.75 | 9.61 | −1.09 | ||||
318.15 | 1.80 | 9.68 | −1.14 | ||||
EB/MAT–Fe3O4 | 303.15 | 0.61 | 8.49 | 5.59 | 5.96 | 20.86 | −0.36 |
308.15 | 0.76 | 8.65 | −0.47 | ||||
313.15 | 0.93 | 8.73 | −0.58 | ||||
318.15 | 1.07 | 8.94 | −0.67 | ||||
EB/JUN–Fe3O4 | 303.15 | 0.59 | 8.48 | 6.29 | 6.54 | 22.70 | −0.35 |
308.15 | 0.70 | 8.59 | −0.43 | ||||
313.15 | 0.90 | 8.71 | −0.56 | ||||
318.15 | 1.07 | 8.95 | −0.66 |
Sample | T (K) | lnKD | lnK2 | E a (kcal mol−1) | ΔH0 (kcal mol−1) | ΔS0 (cal mol−1 K−1) | ΔG0 (kcal mol−1) |
---|---|---|---|---|---|---|---|
MO/ARM–Fe3O4 | 303.15 | 1.14 | 9.03 | 3.27 | 3.31 | 13.24 | −0.69 |
308.15 | 1.28 | 9.17 | −0.78 | ||||
313.15 | 1.33 | 9.21 | −0.83 | ||||
318.15 | 1.41 | 9.30 | −0.89 | ||||
MO/ROS–Fe3O4 | 303.15 | 1.07 | 8.96 | 3.66 | 3.78 | 14.63 | −0.65 |
308.15 | 1.20 | 9.09 | −0.74 | ||||
313.15 | 1.30 | 9.16 | −0.81 | ||||
318.15 | 1.37 | 9.26 | −0.87 | ||||
MO/MAT–Fe3O4 | 303.15 | 0.13 | 8.02 | 6.28 | 6.75 | 22.54 | −0.078 |
308.15 | 0.31 | 8.20 | −0.19 | ||||
313.15 | 0.53 | 8.30 | −0.33 | ||||
318.15 | 0.64 | 8.53 | −0.41 | ||||
MO/JUN–Fe3O4 | 303.15 | −0.27 | 7.62 | 8.45 | 8.67 | 27.27 | +0.16 |
308.15 | −0.13 | 7.76 | +0.08 | ||||
313.15 | +0.12 | 8.00 | −0.07 | ||||
318.15 | +0.39 | 8.28 | −0.24 |
Fig. 11 (a) Plots of lnK2versus 1/T of EB adsorption on Fe3O4 surfaces. (b) Plots of lnKDversus 1/T of EB adsorption on Fe3O4 surfaces. |
Fig. 12 (a) Plots of lnK2versus 1/T of MO adsorption on Fe3O4 surfaces. (b) Plots of lnKDversus 1/T of MO adsorption on Fe3O4 surfaces. |
The activation entropies in all EB/plant–Fe3O4 and MO/plant–Fe3O4 systems are positive, which reveals the affinity of Fe3O4 surfaces for EB and MO molecules. The increasing randomness at the EB/plant–magnetite and MO/plant–magnetite solution interfaces indicates that significant changes in the number of surface active hydroxyl groups occurred in the internal structure of Fe3O4 surfaces. However, activation entropies of EB/JUN–Fe3O4 (22.70 cal mol−1 K−1) and MO/JUN–Fe3O4 (27.27 cal mol−1 K−1) systems are the highest ones, and those of EB/ARM–Fe3O4 (13.03 cal mol−1 K−1) and MO/ARM–Fe3O4 (13.24 cal mol−1 K−1) systems are the lowest ones. This indicates that the changes occurring in the structure of the JUN–Fe3O4 surface are the greatest ones, followed by those of MAT–Fe3O4, then ROS–Fe3O4, and finally ARM–Fe3O4 surfaces.56,57
The activation free energies of EB/ARM–Fe3O4 (−1.09, −1.17, −1.23, and −1.29 kcal mol−1), EB/ROS–Fe3O4 (−0.93, −1.00, −1.09, and −1.14 kcal mol−1), EB/MAT–Fe3O4 (−0.36, −0.47, −0.58, and −0.67 kcal mol−1), and EB/JUN–Fe3O4 (−0.35, −0.43, −0.56, and −0.66 kcal mol−1) systems are negative. However, activation energies of the EB/ARM–Fe3O4 system are more negative than those of EB/ROS–Fe3O4, EB/MAT–Fe3O4 and EB/JUN–Fe3O4 systems, which indicates the feasibility of the EB adsorption process and its spontaneous nature with more EB adsorption on ARM–Fe3O4, then on ROS–Fe3O4, next on MAT–Fe3O4, and finally on JUN–Fe3O4 surfaces.
The activation free energies of MO/ARM–Fe3O4 (−0.69, −0.78, −0.83, and −0.89 kcal mol−1), MO/ROS–Fe3O4 (−0.65, −0.74, −0.81, and −0.87 kcal mol−1), and MO/MAT–Fe3O4 (−0.078, −0.19, −0.33, and −0.41 kcal mol−1) systems are negative. However, activation energies of the MO/ARM–Fe3O4 system are more negative than those of MO/ROS–Fe3O4 and MO/MAT–Fe3O4 systems, which indicates the feasibility of the MO adsorption process and its spontaneous nature with more MO adsorption on ARM–Fe3O4 than on ROS–Fe3O4 surfaces. In the MO/JUN–Fe3O4 system, the values of activation free energy are negative only at 313.15 K and 318.15 K (−0.072 and −0.24 kcal mol−1, respectively), while positive values are found at 303.15 K and 308.15 K (0.16 and 0.079 kcal mol−1, respectively) revealing that activated MO/Fe3O4 complexes are in an excited form in the transition state.56 This leads to the spontaneity of MO adsorption at 313.15 K and 318.15 K.
As presented in Table 8, the found activation energies (Ea) for EB adsorption on ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4, and JUN–Fe3O4 surfaces are respectively: 2.79, 3.21, 5.59, and 6.29 kcal mol−1. Ea is calculated from the slopes of the Arrhenius linear plots ln K2versus 1/T (Fig. 11a). As presented in Table 9, the found activation energies (Ea) for MO adsorption on ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4, and JUN–Fe3O4 surfaces are respectively: 3.27, 3.66, 6.28, and 8.45 kcal mol−1. Ea is calculated from the slopes of Arrhenius linear plots lnK2versus 1/T (Fig. 12a). The found low Ea suggests that EB and MO adsorption processes on all plant–Fe3O4 surfaces proceeded with low energy barriers and can be achieved at relatively low temperatures. As it is known that the activation energy Ea of physical adsorption ranges from 1.2 to 12 kcal mol−1, and from 14.3 to 191 kcal mol−1 for chemical adsorption,58 the adsorption processes of EB and MO on all plant–Fe3O4 are physical in nature.
Fig. 13 (a and b) Temperature effect on EB and MO adsorption yields on plant–Fe3O4 surfaces in the range of 303.15–318.15 K over 20 minutes, respectively. |
Sample | 298.15 K | 303.15 K | 308.15 K | 313.15 K | 318.15 K | |||||
---|---|---|---|---|---|---|---|---|---|---|
q e (mg g−1) | R (%) | q eT (mg g−1) | R T (%) | q eT (mg g−1) | R T (%) | q eT (mg g−1) | R T (%) | q eT (mg g−1) | R T (%) | |
EB/ARM–Fe3O4 | 25.39 | 86.05 | 25.47 | 86.24 | 25.91 | 87.21 | 26.00 | 87.84 | 26.21 | 88.56 |
EB/ROS–Fe3O4 | 22.92 | 77.44 | 24.37 | 82.34 | 24.83 | 83.87 | 25.2 | 85.14 | 25.09 | 85.77 |
EB/MAT–Fe3O4 | 18.03 | 60.91 | 19.44 | 65.66 | 20.24 | 68.38 | 21.12 | 71.35 | 22.19 | 74.95 |
EB/JUN–Fe3O4 | 16.83 | 56.88 | 19.07 | 64.41 | 19.76 | 66.76 | 21.04 | 71.08 | 21.52 | 72.70 |
MO/ARM–Fe3O4 | 20.82 | 70.32 | 21.92 | 74.05 | 22.29 | 75.32 | 22.43 | 75.76 | 22.58 | 76.31 |
MO/ROS–Fe3O4 | 19.80 | 66.88 | 20.59 | 69.55 | 20.93 | 70.72 | 21.65 | 73.15 | 22.05 | 74.50 |
MO/MAT–Fe3O4 | 14.78 | 49.94 | 15.76 | 53.24 | 17.09 | 57.75 | 18.61 | 62.88 | 19.41 | 65.59 |
MO/JUN–Fe3O4 | 11.72 | 39.61 | 12.82 | 43.33 | 13.84 | 46.76 | 15.65 | 52.88 | 17.63 | 59.55 |
The tendency of adsorption capacities and yields on the four magnetite surfaces is the same in EB and MO adsorption processes at 298.15 K and after increasing the temperature from 303.15 to 318.15 K. In EB and MO adsorption processes (298.15 K), the highest adsorption capacities were on ARM–Fe3O4, then on ROS–Fe3O4, next on MAT–Fe3O4, and finally on JUN–Fe3O4 NPs. After the exposure of EB/plant–magnetite and MO/plant–magnetite systems to heat in the temperature range of 303.15–318.15 K for 20 minutes, the order of adsorption capacities was the same.
Table 11 and Fig. 14 show that the adsorption yields and capacities of dyes differed on the four Fe3O4 NPs according to the pH of plant extracts used in magnetite sample synthesis. EB and MO anions were highly adsorbed on the ARM–Fe3O4 surface with achieved adsorption yields and capacities of 86.05%, 25.39 mg g−1 and 70.31%, 20.82 mg g−1, respectively, then on the ROS–Fe3O4 surface with achieved adsorption yields of 77.71%, 22.92 mg g−1 and 66.88%, 19.80 mg g−1, respectively, next on the MAT–Fe3O4 surface with achieved adsorption yields of 61.98%, 18.03 mg g−1 and 49.94%, 14.78 mg g−1, respectively, and finally, on the JUN–Fe3O4 surface where the adsorption yields and capacities of EB and MO achieved values of only 56.88%, 16.83 mg g−1 and 39.61%, 11.73 mg g−1, respectively.When magnetite is immersed in an aqueous acidic solution, it develops its surface charge via the protonation and deprotonation of FeOH sites on its surface according to the following equation:59
FeOH+2 ↔ FeOH0 + H+sol (pka1 = 5.1) | (23) |
Adsorbent | q e (EB) (mg g−1) | R (EB) (%) | q e (MO) (mg g−1) | R (MO) (%) | pH of plant extract |
---|---|---|---|---|---|
ARM–Fe3O4 | 25.39 | 86.05 | 20.82 | 70.31 | 5.25 |
ROS–Fe3O4 | 22.92 | 77.71 | 19.80 | 66.88 | 5.05 |
MAT–Fe3O4 | 18.03 | 61.98 | 14.78 | 49.94 | 4.63 |
JUN–Fe3O4 | 16.83 | 56.88 | 11.73 | 39.61 | 3.69 |
Fig. 14 Adsorption yields of (a) MO and (b) EB dyes on different magnetite surfaces. Error bars represent the standard deviation of three replicates. |
In order to study how the plant extract's acidity impacted preferential attachment of dyes' chromophore and auxochrome groups on synthesized magnetite surfaces, the free chromophore and auxochrome groups that were not attached to magnetite surfaces have been deeply analyzed in all dyes' residual solutions using FTIR spectroscopy, so as to perceive preferential attachment of chromophore and auxochrome groups on each surface and hence determine Brønsted and Lewis acid site densities. Based on this analysis, it was possible to infer Brønsted and Lewis acid site densities on each magnetite surface. For this purpose, after the accomplishment of the adsorption in all experiments, the solid and liquid fractions were separated using a centrifuge. In the next sections, the analysis of residual dye chemistry changes of MO and EB will be studied in detail by comparison between FTIR spectra of MO and EB reference solution chemistry, and the chemistry of their residual solutions.
In the rest of this paper, MO residual solutions will be denoted as MO/ARM–Fe3O4, MO/ROS–Fe3O4, MO/MAT–Fe3O4, and MO/JUN–Fe3O4. Meanwhile, EB residual solutions will be denoted as EB/ARM–Fe3O4, EB/ROS–Fe3O4, EB/MAT–Fe3O4, and EB/JUN–Fe3O4.
Fig. 15 FTIR spectra of (A) the MO reference solution, (B) MO/JUN–Fe3O4, (C) MO/MAT–Fe3O4, (D) MO/ROS–Fe3O4, and (E) MO/ARM–Fe3O4 residual solutions. |
In Fig. 15, spectra (B)–(E) present the FTIR spectra of MO/JUN–Fe3O4, MO/MAT–Fe3O4, MO/ROS–Fe3O4, and MO/ARM–Fe3O4 residual solutions, respectively. The disappearance of certain peaks and the appearance of new ones indicate that the MO/plant–magnetite residual solution chemistry was changed as a result of MO adsorption on magnetite surfaces. The newly appeared peaks correspond to SO−3, , asymmetric CH, phenyl, and the ring skeleton of benzene groups (RSB). More details about the identification of these peaks in MO/plant–magnetite residual solutions will be given in the next paragraphs, for the purpose of investigating the preferential attachment of chromophore and auxochrome groups on all four magnetite surfaces.
The decomposition of the dye is significantly accelerated by the presence of acidic centers at the surface.66
Minor peaks corresponding to NN, CC, and CN groups, linked to the benzene ring, appear in the FTIR spectra of MO/MAT–Fe3O4 and MO/JUN–Fe3O4. Those peaks appearing in the FTIR spectrum of MO/JUN–Fe3O4 are slightly more intense than in that of MO/MAT–Fe3O4, while they do not appear in the FTIR spectra of MO/ROS–Fe3O4 and MO/ARM–Fe3O4 (Fig. 15). The lack of these groups in MO/ROS–Fe3O4 and MO/ARM–Fe3O4 indicates their complete attachment on ROS–Fe3O4 and ARM–Fe3O4 surfaces. However, these groups were almost completely attached on MAT–Fe3O4 and they were less attached on the JUN–Fe3O4 surface.
The analysis of preferential attachment of chromophore groups shows, as summarized in Fig. 16, that:
• On JUN–Fe3O4, all MO chromophore groups were less attached compared to on the three other magnetite NPs.
• On MAT–Fe3O4, NN, CC, and CN groups were almost completely attached while other groups were more attached compared to on JUN–Fe3O4 and less attached compared to on ARM–Fe3O4 and ROS–Fe3O4.
• On ROS–Fe3O4, NN, CC, and CN were completely attached, while other groups were less attached than on ARM–Fe3O4 and more attached than on MAT–Fe3O4 and JUN–Fe3O4.
• On ARM–Fe3O4, NN, CC, and CN were completely attached, whereas phenyldiazonium, phenyl, and benzene groups were more attached on ARM–Fe3O4 than on the other magnetite NPs.
As it is known that chromophore groups prefer to attach to Lewis acid sites, it is possible to infer that the density of Lewis acid sites of ARM–Fe3O4 is the highest one, followed by that of ROS–Fe3O4, then that of MAT–Fe3O4, and finally, that of JUN–Fe3O4.
The functional group attachment analysis results are consistent with MO adsorption yields, being the highest on ARM–Fe3O4 (70.31%), then on ROS–Fe3O4 (66.98%), next on MAT–Fe3O4 (49.94%), and finally on JUN–Fe3O4 (39.61%). This is due to the fact that most of the MO functional groups are chromophores.
Furthermore, FTIR spectra of all four residual solutions (Fig. 15) show that new peaks of asymmetric vibration of CH of CH3 in ionized dimethylamine (DMA) groups appear at around 2342 cm−1,68 and their NCN and stretching bonds at around 2090 cm−1 and 1739.94 cm−1, respectively, with however remarkably different areas. The broadest peak area appears in the FTIR spectrum of MO/ARM–Fe3O4, next in that of MO/ROS–Fe3O4, then in that of MO/MAT–Fe3O4, and finally, the narrowest ones are in the FTIR spectrum of MO/JUN–Fe3O4. This leads to the conclusion that groups were more attached on JUN–Fe3O4, then on MAT–Fe3O4, next on ROS–Fe3O4, and finally, on ARM–Fe3O4.
The analysis of preferential attachment of auxochrome groups shows that:
• On ARM–Fe3O4, all MO auxochrome groups were less attached compared to on the three other magnetite NPs.
• On ROS–Fe3O4, sulphonic acid and dimethylamine groups were more attached compared to on ARM–Fe3O4.
• On MAT–Fe3O4, sulphonic acid groups were completely attached; however, dimethylamine groups were less attached compared to on JUN–Fe3O4.
• On JUN–Fe3O4, sulphonic acid groups were completely attached, and dimethylamine groups were more attached compared to on MAT–Fe3O4.
This leads to the conclusion that the density of Brønsted acid sites of JUN–Fe3O4 is the highest one, followed by that of MAT–Fe3O4, then that of ROS–Fe3O4, and finally, that of ARM–Fe3O4.
Fig. 17 FTIR spectra of (A) the EB reference solution, (B) EB/JUN–Fe3O4, (C) EB/MAT–Fe3O4, (D) EB/ROS–Fe3O4, and (E) EB/ARM–Fe3O4 residual solutions. |
FTIR spectra of EB/JUN–Fe3O4, EB/MAT–Fe3O4, EB/ROS–Fe3O4, and EB/ARM–Fe3O4 are presented in Fig. 17. The disappearance of peaks and the appearance of new peaks indicate that the EB/plant–magnetite residual solution chemistry was changed due to EB adsorption on ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4, and JUN–Fe3O4 surfaces. The newly appeared peaks correspond to not attached (residual) SO−3, phenol, aniline, and phenyl groups, and the ring skeleton of benzene groups. More details about the identification of these peaks in EB/plant–magnetite residual solutions will be given in the next paragraphs, for the purpose of investigating the preferential attachment of the EB chromophore and auxochrome groups on all four magnetite surfaces.
Fig. 17 shows that the peak of the CC bond appears with different peak areas in the FTIR spectra of EB/MAT–Fe3O4 and EB/JUN–Fe3O4 (in the FTIR spectrum of EB/JUN–Fe3O4 it is broader than that of EB/MAT–Fe3O4 and attached on MAT–Fe3O4 more than in JUN–Fe3O4). In contrast, no peak corresponding to the CC bond appears in the FTIR spectra of EB/ROS–Fe3O4 and EB/ARM–Fe3O4 which confirms that CC bonds were completely attached on ROS–Fe3O4 and ARM–Fe3O4 surfaces, and almost completely attached on MAT–Fe3O4; however, these groups were less attached on the JUN–Fe3O4 surface.
Furthermore, the spectrum of EB/JUN–Fe3O4 shows minor peaks of NN, CC, and CN (linked to the benzene ring) groups, whereas no peak corresponding to them appear in other FTIR spectra. This reveals that these groups were completely attached on MAT–Fe3O4, ROS–Fe3O4, and ARM–Fe3O4; however, these groups were less attached on the JUN–Fe3O4 surface.
It is also remarked that new peaks of the CH bond of CH3 of toluene groups appear at around 2343.82 cm−168 in the FTIR spectrum of EB/JUN–Fe3O4 and a minor peak of toluene groups appears in the FTIR spectrum of EB/MAT–Fe3O4. Meanwhile, no peak corresponding to toluene groups appears in the FTIR spectra of EB/ARM–Fe3O4 and EB/ROS–Fe3O4. This reveals that toluene groups were completely attached on ROS–Fe3O4 and ARM–Fe3O4, and almost completely attached on MAT–Fe3O4; however, these groups were less attached on the JUN–Fe3O4 surface.
The analysis of preferential attachment of chromophore groups on magnetite surfaces shows, as summarized in Fig. 18, that:
• On JUN–Fe3O4, all EB chromophore groups were less attached compared to on the three other magnetite NPs.
• On MAT–Fe3O4, NN, CC, and CN were completely attached, whereas CC and toluene groups were almost completely attached while other groups were more attached compared to on JUN–Fe3O4 and less attached compared to on ARM–Fe3O4 and ROS–Fe3O4.
• On ROS–Fe3O4, NN, CC, CN, toluene, and CC were completely attached, while other groups were less attached than on ARM–Fe3O4 and more attached than on MAT–Fe3O4 and JUN–Fe3O4.
• On ARM–Fe3O4, NN, CC, CN, toluene, and CC were completely attached, while phenyldiazonium, phenyl, and benzene groups were more attached than on other magnetite surfaces.
As it is known that chromophore groups prefer to attach to Lewis acid sites, it is possible to infer that the density of Lewis acid sites of ARM–Fe3O4 is the highest one, followed by that of ROS–Fe3O4, then that of MAT–Fe3O4, and finally that of JUN–Fe3O4.
The functional group attachment analysis results are consistent with EB adsorption yields, being the highest on ARM–Fe3O4 (86.05%), then on ROS–Fe3O4 (77.71%), next on MAT–Fe3O4 (61.98%), and finally on JUN–Fe3O4 (56.88%). This is due to the fact that most of the EB functional groups are chromophores.
The analysis of preferential attachment of auxochrome groups shows that:
• On ARM–Fe3O4, all EB auxochrome groups were less attached compared to on the three other magnetite NPs.
• On ROS–Fe3O4, sulphonic acid, phenol, and aniline groups were almost completely attached.
• On MAT–Fe3O4, all auxochrome groups were completely attached.
• On JUN–Fe3O4, all auxochrome groups were completely attached.
This leads to the conclusion that the densities of Brønsted acid sites of JUN–Fe3O4 and MAT–Fe3O4 surfaces are higher than those of ARM–Fe3O4 and ROS–Fe3O4 surfaces.
The analysis of the preferential attachment of chromophore and auxochrome groups in EB and MO adsorption leads to the conclusion that the Lewis acid site density is the highest on ARM–Fe3O4, next on ROS–Fe3O4, then on MAT–Fe3O4, and finally on the JUN–Fe3O4 surface. Moreover, EB and MO adsorption yields and capacities were highest on ARM–Fe3O4 (86.05%, 25.39 mg g−1 and 70.31%, 20.82 mg g−1, respectively), next on ROS–Fe3O4 (77.71%, 22.92 mg g−1 and 66.88%, 19.80 mg g−1), then on MAT–Fe3O4 (61.98%, 18.03 mg g−1 and 49.94%, 14.78 mg g−1), and finally on JUN–Fe3O4 (56.88%, 16.83 mg g−1 and 39.61%, 11.73 mg g−1), respectively. Accordingly, the adsorption yields and Lewis acid site densities varied in the same manner. Seeing that plant extracts used in the synthesis of ARM–Fe3O4, ROS–Fe3O4, MAT–Fe3O4 and –Fe3O4 NPs have respectively pH 5.25, 5.05, 4.63, and 3.69, one can conclude that plant extract pH has a clear effect on the preferential attachment of dye chromophore and auxochrome groups, magnetite nanoparticle acid sites, and adsorption yields. Indeed, the decrease of the mediating plant extract's acidity leads to the increase of Lewis acid site densities and the decrease of Brønsted acid site densities on magnetite NPs and hence an increase in the attachment of chromophore groups and a decrease in the attachment of auxochrome groups of dyes. As most of the MO and EB functional groups are chromophores, the decrease of the mediating plant extract's acidity also leads to an increase in adsorption yields. The remarked difference in adsorption yields of EB and MO on all four magnetite NPs is due to the fact that the ratio of chromophore/auxochrome groups in EB is remarkably greater than that in MO.
Thus, the plant extract's acidity could provide a preconceived idea about the densities of Brønsted and Lewis acid sites of magnetite NPs to be greenly synthesized and therefore about azo dye adsorption yields. Dye adsorption yield can be predicted according to the content of chromophore and auxochrome groups in the azo dye structure.
Sample | pH | EB des. (%) | STD | MO des. (%) | STD |
---|---|---|---|---|---|
JUN–Fe3O4 | 8 | 59.35 | 1.65 | 57.91 | 1.71 |
MAT–Fe3O4 | 8 | 61.35 | 1.83 | 60.38 | 1.69 |
ROS–Fe3O4 | 8 | 64.62 | 1.75 | 61.72 | 1.83 |
ARM–Fe3O4 | 8 | 69.05 | 1.88 | 62.05 | 1.59 |
JUN–Fe3O4 | 9 | 73.25 | 1.67 | 69.45 | 1.62 |
MAT–Fe3O4 | 9 | 74.15 | 1.78 | 72.74 | 1.57 |
ROS–Fe3O4 | 9 | 77.42 | 1.70 | 75.42 | 1.61 |
ARM–Fe3O4 | 9 | 81.15 | 1.62 | 79.05 | 1.85 |
JUN–Fe3O4 | 10 | 80.73 | 1.78 | 78.88 | 1.96 |
MAT–Fe3O4 | 10 | 81.25 | 1.69 | 80.82 | 1.55 |
ROS–Fe3O4 | 10 | 84.44 | 1.65 | 81.84 | 1.78 |
ARM–Fe3O4 | 10 | 89.09 | 1.57 | 84.79 | 1.77 |
JUN–Fe3O4 | 11 | 91.05 | 1.81 | 85.75 | 1.79 |
MAT–Fe3O4 | 11 | 92.19 | 1.68 | 88.11 | 1.74 |
ROS–Fe3O4 | 11 | 95.77 | 1.59 | 89.01 | 1.85 |
ARM–Fe3O4 | 11 | 97.10 | 1.49 | 93.90 | 1.87 |
JUN–Fe3O4 | 12 | 100.0 | 1.89 | 100.0 | 1.71 |
MAT–Fe3O4 | 12 | 100.0 | 1.59 | 100.0 | 1.61 |
ROS–Fe3O4 | 12 | 100.0 | 1.55 | 100.0 | 1.88 |
ARM–Fe3O4 | 12 | 100.0 | 1.54 | 100.0 | 1.63 |
Fig. 19 Stability of magnetite samples in (a) EB and (b) MO adsorption experiments after 3 cycles of reuse. |
Obtained results show that the mediating plant extract's acidity has a clear effect on preferential attachment of dye chromophore and auxochrome groups, magnetite nanoparticle acid sites, and adsorption yields. Indeed, the decrease of plant extract acidity leads to the increase of Lewis acid site densities and the decrease of Brønsted acid site densities on magnetite NPs and hence an increase in the attachment of chromophore dye groups and a decrease in the attachment of auxochrome dye groups. As most of the MO and EB functional groups are chromophores, the decrease of the mediating plant extract's acidity also leads to an increase in adsorption yields.
The linear and non-linear pseudo-first-order and pseudo-second-order kinetics of the adsorption as well as the intra-particle diffusion mechanism have also been analyzed. Obtained results indicated that the adsorption kinetic followed a linear pseudo-first-order kinetic model. Meanwhile, film diffusion was found to be the step that controlled the adsorption mechanism of MO and EB adsorption processes. The thermodynamic studies of EB and MO adsorption processes have been analyzed in the temperature range of 303.15–318.15 K. They reveal the physical and endothermic nature of the adsorption in all cases.
This journal is © The Royal Society of Chemistry 2022 |