3.1. Preparation and Characterization
Herein, the P(SSNa-co-NMAM) was synthesized by free radical polymerization. In the P(SSNa-co-NMAM), SSNa contains sulfonic acid moieties as active sites for cationic dye adsorption (Goswami and Phukan 2017), and the function of NMAM units are thermo-crosslinking points (Yldz et al. 2001). The successful preparation of the P(SSNa-co-NMAM) was demonstrated by the FT-IR and 1H NMR tests. In Figure 1a. the characteristic adsorption bands of sulfonic acid groups at 1007 cm−1 and 1035 cm−1 can be observed on corresponding FTIR spectra. Further, the chemical composition of the synthesized copolymer can be confirmed by 1H NMR spectrum (Figure S1). It suggested that approximately 9.8 mol% of NMAM units in the compolymer after calculating the integral areas between the peak at 6.5 ppm from the SSNa units and the broad peak at 0.6 ∽ 2.2 ppm from the backbone of P(SSNa-co-NMAM), which was close to the NMAM feeding ratio of 10 mol%. These results of 1H NMR further confirmed the target P(SSNa-co-NMAM) was successfully synthesized. Subsequently, a cotton fiber with P(SSNa-co-NMAM)-coating was fabricated by sample thermo-crosslinking. Herein, NMAM was employed as the thermo-crosslinker to anchor P(SSNa-co-NMAM) on the cotton fiberous surface due to N-methylol moieties highly efficient reaction with -OH groups from the cellulose and themselves at high temperature (Figure S2) (Yin et al. 2011).
To explore the chemical structures of as-prepared P(SSNa-co-NMAM) coated cotton fibers, the FTIR measurement was employed. The results were shown in Figure 1a. From the spectrum of PCFs, the adsorption bands at 1497 cm−1 and 1640 cm−1 were attributed to the skeletal vibration of the aromatic ring (C=C) from SSNa units, and the characteristic adsorption bands at 1007 cm−1 and 1035 cm−1 belonged to the sulfonic sodium groups from SSNa units. Meanwhile, the characteristic adsorption band at 833 cm−1 and 1408 cm−1 were assigned to C=O stretching and the C-N stretching vibration of the secondary amide for NMAM units. Therefore, these results imply successful formation of P(SSNa-co-NMAM) polymeric layers on the cotton fibrous surfaces. As expected, the intensity of these characteristic absorption bands from the P(SSNa-co-NMAM) increased when cotton fibers were introduced into the higher concentration coating solution (Figure 1a), further indicating that the coated polymer amount on the cotton fibers increased as polymeric souliton concentration increased. Moreover, the mass change of the PCF before and after treatment were summarized in Table S1, which suggested the coated weight of the P(SSNa-co-NMAM) on the cotton fibers.
Afterward, the surface chemical composition of the PCF was analyzed by the XPS. As shown in Figure 1b, the main peaks at 288.5, 535.5 and 981.5 eV in cotton fibers were ascribed to the binding energy of C 1s, O 1s and O (A), respectively. However, new peaks appeared at 1075.5 eV, 500.6 eV, 233.4 eV, 172.5 eV and 402.5 eV can be observed after immobilization of P(SSNa-co-NMAM) on the cotton fibers, which were assigned to the S 2s, S 2p, Na 1s, Na (A) and N 1, respectively.
Next, the micromorphology of the PCFs was visualized through SEM measurement, and the corresponding SEM images were presented in Figure 2. As shown in Figure 2a-c, the pristine cotton fiber has a smooth surface. However, the P(SSNa-co-NMAM) coated cotton fibrous surfaces could clearly observe polymeric layer (Figure 2d-f). Moreover, the SEM images can verify that as-prepared samples with different loading amounts of the P(SSNa-co-NMAM) could be easily adjusted by the concentration of the polymeric solution (Figure S3). However, the concentration of the polymeric solution was 7%, the intact polymeric layer could be formed among the cotton fibers, which would significantly decrease the specific surface area of the samples. Furthermore, the energy-dispersive spectroscopy (EDS) of the sample was performed. The corresponding EDS spectrum (Figure 2g) and EDS mapping images (Figure 2h-k) were presented. The elemental mapping demonstrated a homogeneous distribution of Na and S elements derived from SSNa units on the PCF, indicating that P(SSNa-co-NMAM) are homogeneously coated on the entire cotton fibrous matrix. Until now, it can be concluded from the FTIR, XPS and SEM results that the P(SSNa-co-NMAM) coated cotton fibers were successfully prepared.
3.2. Adsorption isotherm
Next, the adsorption capacities of the as-prepared PCFs were investigated. As depicted in Figure S5, the adsorption capacity of the PCFs toward cationic dyes significantly increased with increment of the coated mass of P(SSNa-co-NMAM) on the cotton fibers. Notably, the adsorption capacity of the PCF-7 only slightly enhanced compared with that of PCF-5. The reason can be attributed to that some polymeric films were formed on the PCF-7 surface, resulting in decrease of its specific surface area compared to that of PCF-5. As a result, the active sites on the PCF-7 without significantly increased to adsorb dyes compared with that of PCF-5, implying the optimum composition ratio of the PCF-5. Thus, the PCF-5 was selected for the following experiments.
The excellent adsorption capacities are significant for absorbents which affect the practical application of materials. Hence, for investigation of adsorption isotherms of the PCF-5 toward cationic dyes, MEB, RhB, and MG were employed as the cationic dye models, and Figure S6 displays chemical structures of the dyes. As presented in Figure 3a, it could be observed that the adsorption capacity of the PCF-5 toward MEB showed an escalation tendency at the beginning of dye concentration. However, the plateau region appeared when a certain equilibrium concentration was attained. Meanwhile, the adsorption isotherms of PCF-5 toward MEB, RhB, and MG were presented in Figure 3b, which displaed the maximum adsorption capacites of 3976.10 mg/g for MEB, 2879.80 mg/g for RhB and 3071.55 mg/g for MG respectively. The difference of Qm values may be ascribed to the varying chemical structures of dyes resulting in the different interactions between the sulfonate moieties of PCF-5 and dye molecules (Zhan et al. 2021).
Furthermore, the adsorption isotherm is benifical to reveal the mechanism of cationic dyes onto the PCF-5. Herein, the adsorption data were fitted by the Langmuir and Freundlich empirical isotherm models. The Langmuir model proposes a monolayer sorption onto a homogeneous surface and and there is no interaction between the adsorbed particles, while the Freundlich model assumes that multilayer adsorption occurs resulting in the heterogeneous adsorption systems (Song et al. 2020). The linear form of abovementioned two isotherm models are able to be described in the following equations (Zhao et al. 2021a):
The Langmuir model:
In the two equations, qe (mg/g) is absorption capacity (mg/g) at equilibrium and qm (mg/g) is the maximum adsorption, Ce (mg/L) is the concentration of the dye solution at equilibrium; KL (L/mg) is the Langmuir isotherm constant related to the free energy of adsorption; KF (mg/g) and n are the constant of the Freundlich model and heterogeneity factor, respectively. The Langmuir and Freundlich fitting plots of PCF-5 toward MEB, RhB and MG dyes were presented in Figure 3c-d. The correlation coefficients (R2) and model parameters were exhibited in Table 1. By the judgment on the values of the R2, the Langmuir model was more suitable in elucidating the adsorption process of cationic dyes onto the PCF-5 in comparison with Freundlich isotherm model, indicating that monolayer adsorption occurs. Noteworthyly, the dye adsorption capacity of PCF-5 is more competitive compared with numerous similar adsorbents (Table 2), which further confirms that the PCF-5 may be a promising adsorbent andcan be used in wastewater treatment.
Table 1 The results of adsorption isotherms of cationic dyes onto PCF-5
|
Langmuir
|
|
Freundlich
|
dyes
|
KL (L/mg)
|
Qm (mg/g)
|
R2
|
|
KF (L/mg)
|
n
|
R2
|
MEB
|
0.020588
|
3965.10
|
0.99356
|
|
101.83
|
1.793
|
0.86035
|
RhB
|
0.010344
|
2897.80
|
0.9983
|
|
79.13
|
1.955
|
0.90247
|
MG
|
0.006268
|
3074.55
|
0.95121
|
|
6.64
|
0.959
|
0.85878
|
Table 2 The adsorption performance of the PCF-5 comparison of other adsorbents.
Adsorbent
|
absorbate
|
aqmax (mg/g)
|
Ref
|
PP-PDA
|
bMEB
|
434.8
|
(Zhan et al. 2021)
|
Sulfonate-grafted CMPs
|
MEB
|
1650
|
(Zhao et al. 2020)
|
PSS-GR
|
MEB
|
811
|
(Hong et al. 2021)
|
EA-4-6 NFM
|
MEB
|
2257.9
|
(Xu et al. 2019)
|
PSBMA-NaSS
|
MEB
|
760
|
(Xiang et al. 2019)
|
ND-SO3H
|
MEB
|
200.8
|
(Lei et al. 2020)
|
PSSNa/PMMA
|
MEB
|
56
|
(Zhang et al. 2016)
|
PCF-5
|
MEB
|
3976.2
|
this work
|
QPVA/TEOS hybrid membrane
|
cRhB
|
34.16
|
(Zhang et al. 2017)
|
H2O2-modified OBDCA
|
RhB
|
50
|
(Zhao et al. 2021b)
|
PCF-5
|
RhB
|
2879.8
|
this work
|
Modified Irvingia gabonensis
|
dMG
|
250
|
(Abdi et al. 2019)
|
Bentonite
|
MG
|
178.6
|
(Bulut et al. 2008)
|
PCF-5
|
MG
|
3071.6
|
this work
|
|
|
|
|
|
* aqmax is the maximum adsorption capacity. Cationic dye: bMEB, methylene blue; cRhB, rhodamine B; dMG, malachite green.
3.3. Adsorption kinetics
In order to better inquire into the adsorption behaviors of the PCF-5, the effect of contact time on the adsorption of cationic dyes on absorbent was investigated. In Figure 4a, it was clearly observed that the adsorption equilibrium time of dyes increased with the increment of molecular weight of dyes. This result can be explained that dye with higher molecular weight has weaker thermal motion ability and greater steric hindrance, thus it is not easy to reach the adsorption site (Hussain et al. 2022). Generally, the PCF-5 toward cationic dyes exhibited rapid adsorption rate, especially in the first 10 min, due to the presence of abundant active adsorption sites, and then gradually reached to adsorption equilibriums within 50 min. Selecting MEB as an example, a rapid adsorption rate of the PCF-5 was verified by the UV−vis spectra. As we can see from Figure 4b, as the prolongation of adsorption time the absorbance intensity rapidly disappeared, suggesting the rapidly decreased MEB concentrations in the solution within the short time, the corresponding optical picture also implying this phenomenon (Figure 4c).
Adsorption kinetic studies provide important information to reveal the mechanism of the dye adsorption process. Herein, the adsorption processes of PCF-5 toward MEB, RhB and MG were descired via the pseudo-first-order (Eq. (2)) and pseudo-second-order (Eq. (3)) kinetic models, respectively, as expressed in the following equations (Yang et al. 2021a):
The pseudo-first-order model:
Where qe is the amount of adsorbed dye (mg/g) at equilibrium state; qt is the adsorption amount at time t (min); k1 (min−1) and k2 (g/mg/min) denote the rate constant of the pseudo-first-order and the pseudo-second-order adsorption kinetics, respectively. The related kinetic model parameters were listed in Table 3. It could be found that the correlation coefficients (R2) of the pseudo-first-order kinetic model were very closer 1 than that of the the pseudo-second-order model. Moreover, qe,cal calculated by pseudo-second-order model was much closer to the experimental qe,exp as compared with the pseudo-first-order model. These results proved that the adsorption toward MEB of the PCF-5 obeys the pseudo-second-order model. Hence, these kinetic results verified that the chemical interaction between the cationic dye molecules and the sulfonic acid groups of the PCF-5 promoted the adsorption process.
Table 3 The parameters of kinetics for adsorption of cationic dyes onto PCF-5
|
Pseudo-first-order
|
|
Pseudo-second-order
|
dyes
|
k1(min-1)
|
qe1, cal (mg/g)
|
R2
|
k2(g/mg min-1)
|
qe2, cal (mg/g)
|
R2
|
MEB
|
0.1107
|
76.70
|
0.99283
|
|
0.001453
|
198.98
|
0.8768
|
RhB
|
0.1223
|
101.26
|
0.99699
|
|
0.00072
|
166.85
|
0.8337
|
MG
|
0.0785
|
93.22
|
0.99266
|
|
0.000459
|
171.37
|
0.9627
|
3.4. Effect of pH value
The pH value of the water sample acts a very important role in this adsorption system. The changes of the ionic or neutrality state of the target compounds and the charge of the adsorbent with pH would promote or inhibit the interaction between the adsorbates and the adsorbents, and thus affect the adsorption performance (Zhang et al. 2013; Zhao et al. 2015). In Figure 5a, the removal rate of the PCF-5 slightly decreased at pH=2 that because of the protonation of the sulfonic acid group in a highly acidic environment. However, the PCF-5 presented an insignificant effect of pH=2-11 on the adsorption performance. These results demonstrated that the PCF-5 could maintain stable adsorption capabilityin a wide range of pH.
3.5. Effect of ionic strength
Ions can be found commonly in industry wastewater including textile and aquaculture wastewater. Therefore, the existence of salts in wastewater may affect the ionic strength between the adsorbent with the dye molecules (Yang et al. 2021b). Thus, the effect of ions strength was examined by addition of NaCl into the MEB solution. The result shown that the removal efficiency of PCF-5 was significantly inhibited when the salt concentration increased from 0 to 0.6 mol/L (Figure 5b). This phenomenon was probably attributed to the salt ions (Na+) competition with MEB molecules to occupy the active adsorptive site with sulfonic acid groups from the PCF-5. Hence,the increase of salt concentration in MEB molecules will lead to the enhancement of shielding effect of salt ions on positively charged MEB molecules, thus reducing the removal efficiency of adsorbent.. Above phenomenon can be concluded that electrostatic interaction is a significant derive forcein the adsorption process of cationic dyes onto the PCFs.
3.6. Effect of Temperature and Adsorption Thermodynamic Analysis
Another important parameter in adsorption process is the efficacy of temperature on the adsorption rate. Herein, the effect of temperature on the adsorption of dyes on the PCF-5 was investigated under different temperatures (298, 308, and 318 K). As depicted in Figure 6a, the removal rate of PCF-5 toward cationic dyes increased with the increment of temperature. It would be considered that the increment of the temperature facilitates the movement of cationic dye molecules and provides additional energy to enhance the interactions between the cationic dye molecules and the adsorption sites of the PCF-5.
To explore the spontaneous completion of the adsorption process and further to lookinto the adsorption mechanism of the PCF-5. The thermodynamic parameters of dyes adsorption onto PCF-5, including Gibbs free energy (ΔG0), ntropy (ΔS0), and enthalpy (ΔH0), were calculated by the following equations (Lei et al. 2020):
where R (8.314 J mol−1 K−1) is the universal gas constant ,T (K) and Kc (dimensionless) are the absolute temperature and the adsorption equilibrium constant, respectively. ΔG0 could be figured out by Eq. (3), whereas the ΔH0 and ΔS0 are caculated from the slope and intercept of the plot of lnKc versus 1/T (Figure 6b), respectively. The corresponding parameters were listed in Table 4. The negative values of ΔG° confirmed that the adsorption process toward MEB, RhB, and MG onto the PCF-5 was spontaneous nature, indicating that the efficiency of adsorption is more desirable at higher temperatures. Additionally, considering the positive values of ΔH° for MEB onto PCF-5 revealed that an endothermic reaction was in the adsorption process (Far et al. 2020). Furthermore, the fact that the positive value of ΔS° also revealed a randomness increase on the solid-liquid interface between PCF-5 and dye solution. These analyses were consistent with the result that higher temperatures could promote the sorption process of cationic dyes on the PCF-5.
Table 4. Thermodynamic parameters for the adsorption of PCF-5
dyes
|
T(K)
|
ΔG° (KJ/mol)
|
ΔH° (KJ/mol)
|
ΔS° (KJ/mol K)
|
MEB
|
298
|
-8.207
|
0.685
|
0.062
|
308
|
-1.187
|
318
|
-1.421
|
RhB
|
298
|
-6.025
|
308
|
-7.752
|
0.288
|
0.035
|
318
|
-8.903
|
MG
|
298
|
-3.607
|
0.235
|
0.024
|
308
|
-4.886
|
318
|
-5.698
|
3.7. Selective Adsorption and Dynamic Filtration Performance
Based on the negatively-charged property owing to the sulfonic acid groups (-SO3H), the as-prepared PCF-5 was supposed to be beneficial to adsorb cationic dyes due to the electrostatic interaction (Zhou et al. 2018; Zheng et al. 2020). Three cationic dyes (MEB, MG and RhB) and three anionic dyes (MO, CR and NGB) were selected to study the adsorption performance of PCF-5, and Figure S6 shown their corresponding structures. The PCF-5 exhibited an excellent adsorption capacity for MEB, MG and RhB dyes, compared with that of MO, CR and NGB dyes (Figure S7). Based on these conclusions, we further explored the selective adsorption performance of the PCF. As observed from Figure 7a, the color of the MEB/MO mixture solution switched from dark green to yellow after being fully adsorbed by PCF-5. Moreover, the UV–vis spectra suggested that the peak at 664 nm of MEB almost disappeared after the adsorption, while the peaks at 464 nm of MO was only a slight decrease, verifying that MEB was selectively adsorbed from the MEB/MO mixture dyes (Figure 7b).
Further, the PCF-5 was expected to have a highly efficient wastewater purification through the filtration process. Therefore, the dynamic filtration performance of the PCF-5 toward dyes was investigated. As shown in Figure S8, the 20 mL of dark blue MEB solution can be purified by PCF-5 membrane with a gravity-driven force. Besides, the dynamic selective separation measurement of the PCF-5 toward mixed dyes can also be carried out. As we can see, MEB could be successful separated by PCF-5 from MEB/MO mixture dyes during the filtration−separation process (Figure 7c). Meanwhile, the completely disappearing of the adsorption peak from MEB in UV−vis spectrum (Figure 7d) after dynamic filtration further proved that the cationic dyes can be highly effective adsorbed onto the PCF-5 from the mixture under dynamic state owing to the high adsorption efficiency of the PCF-5. Therefore, aforementioned adsorption experiments demonstrated that the PCF-5 possessed outstanding adsorption capacity toward cationic dyes.
3.8. Adsorption mechanism
To gain a deeper understanding on the adsorption behavior of the PCFs toward cationic dyes, adsorption mechanism is discussed in detail. It is not only for potential practical application but also for further exploitation of advanced adsorbents.
As mentioned above, the adsorption process of the PCFs conformed to the Langmuir model and pseudo-second-order kinetic model indicated that was considered to greatly involve the chemical interaction between cationic dyes and PCFs. Taking MEB as an example, the potential interactions between PCFs and MEB were comprehensively investigated. Firstly, the effect of ionic strength of the solution revealed that the adsorption capacity of the PCFs toward cationic dyes significantly decreased with increasing ionic strength, implying that the main driving force of MEB adsorption onto PCFs is the electrostatic interaction (Xu et al. 2018). Furthermore, the π–π interaction might have a role in the MEB adsorption since both the MEB molecules and the PCFs have the aromatic rings (Lei et al. 2020). Thus, these possible interactions between the PCFs and cationic dyes were illustrated in Figure 8. In short, the cationic dyes adsorption of PCFs could be summarized that the major mechanism for the adsorption of dyes is electrostatic interaction, while the π–π interaction also contribution of another adsorption mechanism.
3.9. Adsorption performance in simulated dye effluent
The practical potentials of as-proposed PCF-5 was evaluated for treating a simulated dye effluent with containing multiple components (Chen et al. 2020). Table 5 listed the ingredients of the simulated dye effluent. The adsorption performance of PCF-5 on simulated dye effluent was demonstrated by the UV-vis spectra and corresponding photographs (Figure 9). In Figure 9a, the removal efficiency of PCF-5 could reach to 99.9% in the simulated dye effluent. Furthermore, Figure 9b shown that the original color of the simulated dye effluent changes from purple blue to transparency after treatment by the PCF-5. These results can further manifest that the PCF-5 showed excellent adsorption performance in contaminated wastewater remediation.
Table 5 The composition of the simulated dye effluent
Compound
|
λmax (nm)
|
|
Concentration (mg L−1)
|
MEB
|
664
|
|
20
|
RhB
|
554
|
|
20
|
MG
|
618
|
|
20
|
NaCl
|
|
10
|
K2CO3
|
|
10
|
K2HPO4
|
|
10
|
CaCl2
|
|
10
|
3.10. Recyclability
There is no doubt that the recyclability of adsorbents which is directly related to the cost-effectiveness of an adsorption process is also important indicators for its practical applications. In this work, the recyclability of the PCF-5 was evaluated via using successive cycles of adsorption-desorption. As presented in Figure 10, PCF-5 exhibited a quite stable property with a nearly constant removal efficiency during the recycling experiments. Its removal efficiency still maintained above 90% after 5 successive cycles. Meanwhile, the SEM image of the regenerated PCF-5 was presented in Figure S9a. It displayed that the cotton fibrous surfaces still coated by the P(SSNa-co-NMAM) copolymer well even after 5 successive cycles. Moreover, the S, Na and N peaks of the regenerated PCF-5 still can be obviously observed in the XPS spectra (Figure S9b). Hence, from the results of the cyclic adsorption experiments and the characterization of regenerated PCF-5, it can be concluded that the PCF-5 possessed outstanding durable feature.