Next Article in Journal
Effect of Palm Oil–Carnauba Wax Oleogel That Processed with Ultrasonication on the Physicochemical Properties of Salted Duck Egg White Fortified Instant Noodles
Next Article in Special Issue
Macroalgal-Derived Alginate Soil Amendments for Water Retention, Nutrient Release Rate Reduction, and Soil pH Control
Previous Article in Journal
Biological Thermal Performance of Organic and Inorganic Aerogels as Patches for Photothermal Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of Sustainable Tannin/Sodium Alginate Composite Hydrogel Beads for Efficient Removal of Methylene Blue

Hubei Key Laboratory of Radiation Chemistry and Functional Materials, School of Nuclear Technology and Chemistry & Biology, Hubei University of Science and Technology, Xianning 437100, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Gels 2022, 8(8), 486; https://doi.org/10.3390/gels8080486
Submission received: 9 July 2022 / Revised: 26 July 2022 / Accepted: 1 August 2022 / Published: 3 August 2022
(This article belongs to the Special Issue Alginate-Based Gels: Preparation, Characterization and Application)

Abstract

:
To meet the requirement of sustainable development, bio-based adsorbents were developed for the removal of dye contaminant. To improve the adsorption capacity of pure sodium alginate (SA) adsorbent for the removal of methylene blue (MB), aromatic bio-based tannin (Tan) was incorporated through the cross-linking with calcium ion. The obtained Tan/SA composite hydrogel beads were characterized with SEM, FTIR and TG, demonstrating that millimeter-sized beads were obtained through calcium cross-linking with enhanced thermal stability. The maximum capacity (247.2 mg/g) at optimal condition (pH = 12, T = 45 °C) was obtained for the 40%Tan/SA adsorbents, with a removal efficiency of 82.4%. This can be ascribed to the electrostatic attraction between SA and MB, as well as the formation of π–π stacking between Tan and MB. The adsorption process for MB is endothermic, and chemical adsorption, the removal efficiency was exceeded 90% after five cycles.

Graphical Abstract

1. Introduction

Synthetic dyes have been widely used in textile industries, causing serious pollution in water resources. Once the synthetic dyes enter the human body, it can endanger people’s health and cause a variety of diseases [1,2]. The synthetic dyes are classified into three types, namely cationic dye, anionic dye and non-ionic dye, according to the charge of the molecules. Among these dyes, methylene blue (MB) is a typical cationic dye which has been widely used in the field of textile, paper, coating and printing. It can easily combine with oppositely charged cell membranes, leading to a high hazard for organisms. In addition, the MB molecules possess a very high color value, which can be clearly observed even at relatively low concentrations (<1 mg/L) [3,4]. Therefore, the removal of MB from wastewater is an important issue.
To meet the requirements of sustainable development, many bio-based adsorbents, especially for composite hydrogel were fabricated for the adsorption of MB in wastewater. Activated lignin-chitosan extruded pellets were reported for the removal of MB, the adsorption results fitted well with the Langmuir isotherm, while the maximum adsorption capacity is 36.25 mg/g [5]. Chitosan/sepiolite composite hydrogel were prepared with cross-linking using epichlorohydrin, the CS50SP50 was selected as the best adsorbent for the adsorption of MB, and the monolayer adsorption capacities of this composite is 40.986 mg/g [6]. Sodium alginate/soybean extract composite hydrogel beads were reported for the removal of MB by Viscusi et al. [7], the adsorption data can be described with a second-order model, and the maximum adsorption capacity is 49 mg/g. Lu et al. reported magnetic activated carbon/sodium alginate composite hydrogel beads for the adsorption of MB [8], and the maximum adsorption capacity is 222.3 mg/g with initial concentration of 700 mg/L, with a removal efficiency of 31.8%. Alginate/activated carbon composite hydrogel beads were also reported for the removal of MB [9], and a higher adsorption capacity (730 mg/L) was obtained with high initial concentration of 1000 mg/L at 40 °C, with a removal efficiency of 73%. Though different bio-based adsorbents were developed for the removal of MB, the adsorption capacity or the removal efficiency need significant improvement.
The intermolecular interaction between the adsorbents and the dye is an important factor to improve the adsorption performance. The commercial dyes are charged molecules with aromatic structures, for example, MB is a positive dibenzothiazine derivative. Based on its structure feature, many different adsorbents were designed and fabricated. Due to its positive charge of MB, electrostatic attraction was taken into account, firstly in the adsorption design. Many anionic polymers were prepared for the removal of MB, including sodium polystyrene sulfonate [10], polyacrylic acid [11] and bio-based sodium alginate [7,8,9]. As π-π stacking interaction can form between dyes and the compound containing an aromatic structure, the introduction of an aromatic structure was taken into account in the adsorbent design. In order to improve the adsorption performance, bio-based aromatic lignin [12] and tannin [13] were introduced into the polymer matrix to enhance the adsorption capacity.
The composites consisting of bio-based tannin (Tan) and sodium alginate (SA) can not only meet the requirement of sustainable development, but also lower the cost and extend the application of the production. Tan/SA films were reported by Kaczmarek et al. [14], and improved physicochemical properties were observed for the composite films. In addition, the films showed antioxidant properties, which have a promising application in traditional packaging material. Larosa et al. reported the characterization of Tan/SA beads by morphology, thermal stability, FTIR and XRD [15], and the non-crystalline property makes them suitable as carriers for possible controlled release, and they could be used in food applications to improve tea quality or clarify juices. The enhanced drug encapsulation and extended release profile of SA with the incorporation of Tan was reported by Abulateefeh et al. [16]; the incorporation of Tan led to a more than fourfold increase in drug encapsulation efficiency and reduced burst drug release from 44% to around 10% within the first 30 min. Though the bio-based Tan/SA composite was prepared and investigated, there was no report for the removal of MB with Tan/SA beads. Therefore, Tan/SA beads were fabricated by cross-linking with calcium ion and used for the removal of MB (Figure 1); the adsorption of MB can be realized with electrostatic attraction originated from SA and can be improved through π-π stacking interaction with incorporated aromatic Tan. The obtained Tan/SA beads were characterized with FTIR, SEM and TG, with the influence of experimental parameters, and the thermodynamic, kinetic and regeneration were also investigated.

2. Results and Discussion

2.1. Characterization of Tan/SA Beads

2.1.1. Morphology Analysis

The photos and surface morphology of pure SA beads and 40% Tan/SA beads are shown in Figure 2a–f. The pure SA beads exhibit colorless transparency, while the color changed to light yellow for the 40% Tan/SA beads. The obtained adsorbents are spherical in shape and show a uniform size distribution, with a diameter of about 2 mm. The prepared composites are millimeter-sized beads, similar to the commercial adsorption resin, which is of benefit for the application in industry [17,18]. The pure SA beads present a smooth surface, while the 40% Tan/SA beads show a rough surface with many folds. The introduction of Tan into the SA matrix enhances the surface area of the obtained adsorbents, which will improve the adsorption capacity [12].

2.1.2. Structure Analysis

As shown in Figure 3, the structure of SA, Tan and 40% Tan/SA beads was revealed by FTIR. In the spectra of SA, the vibrations of –OH were confirmed with the broad peak from 3000 to 3600 cm−1, the presence of stretching vibration (2926 cm−1) and bending vibration (1415 cm−1) revealing the existence of an aliphatic C–H bond. The peak around 1628 cm−1 can be ascribed to the carboxylate ion, and the peaks at 1030 and 950 cm−1 can be assigned to the C–O vibrations [14]. The spectrum of Tan shows a broad band from 3100 to 3500 cm−1, which corresponds to the stretching of –OH in the phenolic groups. In addition, strong and sharp peaks at 1711 and 1608 cm−1 are attributed to the carbonyl stretching of ester groups and aromatic C–O stretching vibrations. The peaks at 1532 and 1452 cm−1 can be attributed to the C–C stretching for the aromatic skeleton [15], which is of benefit for the formation of π–π stacking with aromatic MB in the adsorption, leading to the enhancement of adsorption capacity. For the 40% Tan/SA beads, all the peaks of SA and Tan can be observed, and the peak of 950 cm−1 for the SA disappeared. In addition, the carbonyl stretching peak shifts from 1628 to 1636 cm−1, revealing the coordinate nature of calcium-bonding; thus, the Tan/SA beads were successfully fabricated with the calcium ion cross-linking [14,15].

2.1.3. Thermal Stability Analysis

The stability of the hydrogel beads was investigated, and the SEM images of the pure SA hydrogel beads and Tan/SA-40% hydrogel beads after heat treatment is presented in Figure 4. Under the condition of 50 °C, after shaking for 48 h at 200 rpm, the morphology of pure SA hydrogel beads and Tan/SA-40% hydrogel beads showed no significant change, demonstrating that the hydrogel beads have a good stability in an aqueous solution, which will be stable under the adsorption experimental condition.
The TG curves of Tan, SA and 40% Tan/SA beads are shown in Figure 5. The Tan exhibits a three-stage degradation process; there is a loss of adsorbed water (about 7.5%) before 180 °C. The decomposition of linkage between flavonoid units is in the temperature range of 180–320 °C, with a weight loss of 33%. The pyrolytic degradation of flavonoid units occurs after 320 °C, and the residue char is 22.8% at 700 °C [19,20]. The TGA curves of SA gave four mass loss before 700 °C. The first stage, which is approximately 7.2% before 120 °C, is attributed to the evaporation of absorbed water. In the second stage, the mass loss of about 6.1% between 150 and 220 °C could be due to the decarboxylation of SA released carbon dioxide. The third stage mass lose between 220 and 450 °C is assigned to the thermal decomposition of the aminopropyl groups and bisaldehyde. The last stage indicates that the organic species have been completely decomposed, and the residue char is 38.6% at 700 °C [8]. The 40% Tan/SA beads showed a similar degradation behavior to pure SA, and improved thermal stability was obtained for the composite beads. This can be ascribed to the rigid structure of Tan molecules, which is more stable than the flexible chain of SA [8,12].

2.2. Adsorption Behavior of Tan/SA Hydrogel Beads

The MB is positive charged dye with an aromatic structure. As mentioned in the structure analysis (Figure 3), the obtained Tan/SA beads possess a carboxyl group with a negative charge and aromatic ring (Tan). The MB can be adsorbed onto the Tan/SA beads through the combination of electrostatic attraction, π-π stacking, according to the literature [10,11,12,13]. The fabricated Tan/SA beads were used for the removal of MB, under different adsorption conditions.
As shown in Figure 6, the removal efficiency for the 40% Tan/SA beads under different pH values were investigated first. The adsorption behavior of Tan/SA beads show a pH dependence toward MB, demonstrating that the electrostatic attraction plays an important role in the adsorption process. With the increase in the pH value, the removal efficiency increased obviously, and the highest R (88.46%) was observed at pH = 12. This can be ascribed to the disadvantage of high H+ concentration; the H+ has a competitive adsorption with positive MB [10]. Though a high pH value is of benefit for the enhancement of removal efficiency, taking the protection of the environment into account, adsorption at a higher pH (>12) was not conducted.
The effect of temperature on the removal efficiency was also investigated (Figure 7). With the increase in the solution temperature, the removal efficiency increased dramatically, and the highest R (84.62%) was obtained at 45 °C. This phenomenon revealed that the adsorption is an endothermic process, which may be assigned to the increase in the mobility of MB molecules and the number of active sites of Tan/SA beads [21]. Though high temperature is of benefit for the enhancement of removal efficiency, taking the energy consumption into account, adsorption at a higher temperature (>45 °C) was not conducted.
The thermodynamic parameters, including the equilibrium constant K0, the change of free energy ΔG0, entropy ΔS0 and enthalpy ΔH0 can be calculated using the following equations [12,22,23]:
lnKo = ln(Qe/Ce)
ΔG0 = −RT ln Ko
Rln Ko = −ΔH0/T + ΔS0
The Qe is adsorption capacity (mg/g), Ce is equilibrium concentration, R is the universal gas constant (8.314 J/mol·K) and T is the Kelvin temperature. The ΔH0 and ΔS0 are calculated from the linear plot of ln K0 versus 1/T.
The calculated thermodynamic parameters are listed in Table 1. With the increase in the temperature, the ΔG0 decreased gradually, illustrating that high temperature can accelerate the adsorption process, which is consistent with the results mentioned above. When the temperature is no more than 298 K, the value of ΔG0 is positive, indicating that the adsorption cannot proceed spontaneously. With the increase in temperature, the value of ΔG0 changed to negative, revealing that the adsorption can proceed spontaneously at a high temperature. The ΔH0 and ΔS0 calculated from the slope and intercept are 80.80 kJ/mol and 0.27 kJ/mol·K, respectively. The positive value of ΔH0 demonstrated an endothermic feature of the MB adsorption again, and the positive ΔS0 values showed the good affinity of Tan/SA beads for MB during the adsorption process [24,25,26].
Under the optimal conditions (pH = 12, T = 45 °C), the effect of Tan content on the adsorption capacity was investigated (Figure 8). With the incorporation of Tan, the adsorption capacity of MB enhanced obviously. The highest adsorption capacity (247.2 mg/g) was observed for the 40% Tan/SA beads, with a removal efficiency of 82.4%. In comparison with pure SA beads, the maximum adsorption capacity increased by 264%. This can be ascribed to the formation of strong interaction of π-π stacking, between the aromatic tannin and the MB. With a further increase in the tannin content, the adsorption capacity decreased, which is due to the hindered effect of excessive rigid tannin for the diffusion of MB molecules [12,27].

2.3. Adsorption Kinetics of Tan/SA Hydrogel Beads

As shown in Figure 9, the influence of concentration with the increase in contact time (within 4 h) on adsorption was investigated. The adsorption rates are fast with different initial concentrations (60, 80, 100 mg/g); with the increase in the initial concentration of MB, the equilibrium time increased from 2 to 3 h, revealing that the adsorption of MB is dependent on the concentration. This is because it takes a long time to diffuse into the internal surface for MB molecules, after the saturation adsorption on the external surface of the Tan/SA beads [28,29].
The adsorption kinetic can reveal the adsorption efficiency and determine the potential application of the obtained bio-based Tan/SA composite adsorbents. To give a deep insight into the adsorption process and adsorption mechanism for MB adsorption onto Tan/SA beads, pseudo-first-order, pseudo-second-order and Elovich models were used to evaluate the adsorption kinetic (Equations (4)–(6)) [30,31,32,33,34].
Pseudo-first-order:
log ( q e q t ) = log q e k 1 t
Pseudo-second-order:
t q t = 1 k 2 q e 2 + t q e
Elovich model:
q t = ln α β β + 1 β ln t
where qe and qe are the adsorption capacity at equilibrium and at time t, respectively, k1 and k2 are the rate constant, α is the initial adsorption rate and β is the desorption constant.
The fitting curves of MB adsorption with the three kinetic models are shown in Figure 10, and the parameters calculated from these curves are listed in Table 2. For the equilibrium adsorption capacity at different initial concentrations, the experimental value is very close to the calculated one for the pseudo-second-order model. The correlation coefficient R2 (0.9999) shows that this model fits the experiment data better than that of the pseudo-first-order and Elovich model, demonstrating that the pseudo-second-order is predominant in the adsorption process, and the chemical adsorption is the rate-limiting step for the MB molecules’ adsorption [10,12,35].

2.4. Regeneration Behavior of Tan/SA Hydrogel Beads

Recyclability is an important factor in evaluating the performance of the obtained bio-based Tan/SA absorbents; therefore, the prepared Tan/SA beads were evaluated by the adsorption/desorption experiment. The adsorption experiment was conducted with an MB concentration of 10 mg/L at pH = 12 and T = 45 °C for 4 h. The desorption of MB molecules from Tan/SA was achieved with 1 M HCl aqueous solution in an incubator shaker and washed with distilled water several times. As shown in Figure 11, the removal efficiency of MB decreased slightly, and the removal efficiency still exceeds 90% after five cycles. This result revealed that the adsorption ability was not changed obviously after regeneration, indicating that the Tan/SA adsorbents have a good reusability [10,28].

2.5. Comparative Analysis of MB Adsorption

The adsorption capacity, removal efficiency and regeneration of the obtained Tan/SA beads was compared with similar bio-based adsorbents in reported literature. As shown in Table 3, there was no regeneration data for most of the reported bio-based adsorbents. Some of the reported adsorbents show very high adsorption capacity at much higher initial concentrations, but the removal efficiency is relatively low. The adsorption capacity or removal efficiency of the Tan/SA beads was better than most of the similar bio-based adsorbents. The improved adsorption performance of the obtained bio-based adsorbents can be ascribed to the incorporation of aromatic Tan, which can form π-π stacking interaction with aromatic MB in the adsorption [12,13].

3. Conclusions

Sustainable millimeter-sized Tan/SA composite hydrogel beads were successfully synthesized for the efficient removal of MB, which will benefit environment protection, sustainable development and industrial application. The introduction of aromatic Tan into the SA matrix can not only enhance the thermal stability of the prepared beads, but also improve the adsorption performance through the formation of π-π stacking. The maximum capacity at optimal condition was obtained for the 40%Tan/SA adsorbents, with a removal efficiency of 82.4%. The positive of ΔH0 demonstrated an endothermic feature of the MB adsorption, and the calculated kinetic parameters revealed that the chemical adsorption is the rate-limiting step for the MB molecules’ adsorption. The removal efficiency of MB still exceeds 90% after five cycles, which indicated that the Tan/SA adsorbents have a good regeneration ability. The fabricated bio-based Tan/SA adsorbents have a promising application for the removal of MB in industry. Future work will focus on the selectivity, continuous column adsorption and scale-up of this kind of bio-based adsorbent.

4. Materials and Methods

4.1. Materials

All reagents were purchased from commercial suppliers of analytical grade reagent and used without further purification. Tannin (Tan, 98%, Mw = 1700 g/mol), sodium alginate (SA, 200 mPa·s), CaCl2 and methylene blue (MB, 98%) were purchased from HWRK chemical Co. Ltd. (Beijing, China). Distilled water was used throughout the experiments for solution preparation.

4.2. Preparation of Bio-Based Tan/SA Hydrogel Beads

The bio-based Tan/SA beads were prepared as follows [14,15,16]: the Tan was dispersed into 30 mL water using a bath sonicator. Then, 0.6 g SA was added to the dispersion under stirring until complete dissolution at 60 °C, the weight percentage of Tan to SA is 0, 20, 40, 80, 120 wt%). The obtained mixture was dropwised into 5 wt% CaCl2 solution at room temperature without stirring, the cross-linking reaction kept for 24 h. Tan/SA composite beads were obtained with filtration, washed three times with distilled water. The obtained hydrogel beads were dried at 60 °C with a vacuum drying oven to a constant weight, the percentage of adsorbed water for the hydrogel beads containing 0, 20, 40, 80, 120 wt% Tan is 94.56, 93.88, 93.56, 93.12, 92.33%.

4.3. Characterization

Morphology of the obtained beads were revealed by a scanning electron microscopy under 10 kV (SEM, VEGA-3, Tescan, Czech Republic). The structure of the beads was demonstrated with Fourier transform infrared (FTIR, Avatar 360 Nicolet instrument). The stability of hydrogel beads was investigated in water under 50 °C, with a bath shaker at 200 rpm. The thermogravimetry (TG) analysis was conducted via a TG 209F3 instrument (NETZSCH Scientific Instruments Ltd., Shanghai, China) under N2 atmosphere with a heating rate of 10 °C/min−1. The absorbance of the MB solution was tested with a S 3100 spectrophotometer (Mapada Instruments Co. Ltd., Shanghai, China).

4.4. Adsorption Experiments

The adsorption of MB from aqueous solution onto Tan/SA beads were systematically evaluated. Then, 20 mg dried Tan/SA beads were added into 30 mL of MB solution, and the solution placed in an isothermal water bath shaker at 150 rpm. The pH of the solution was adjusted with 0.1 M HCl and NaOH solution. The concentration of MB was determined by using a UV-vis spectroscopy at 664 nm. The adsorption capacity Q (mg/g) and removal efficiency R (%) were calculated as follows:
Q = (C0Ct)V/m
R = (C0Ct) × 100/C0
C0 (mg/L) and Ct (mg/L) are the initial concentration and remaining concentration of MB, m (g) is the weight of Tan/SA beads and V (L) is the volume of the solution.

Author Contributions

Experiment design, J.G. and T.C.; sample preparation, X.H., Z.W. and Z.L.; structure characterization, X.H., Y.Z. and Z.L.; property analysis, Z.W., Y.Z. and G.H.; writing—original draft, J.G.; review and editing, T.C. and G.H.; supervision, T.C.; project administration, J.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (51903080), Innovation and Entrepreneurship Training Program for College Students of Hubei Province (S202210927005), Innovation Team of Hubei University of Science and Technology (2022T03) and Scientific Research Foundation of Hubei University of Science and Technology (BK202003, 2021ZX15).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghahehe, F.S.; Taghizadeh, M.; Taghizadeh, A.; Hatati, B.; Mahmoodi, N.M.; Parastar, S. Clean synthesis of rock candy-like metal–organic framework biocomposite for toxic contaminants remediation. Environ. Technol. Innov. 2021, 23, 101747. [Google Scholar] [CrossRef]
  2. Mahmoodi, N.M.; Taghizadeh, M.; Taghizadeh, A. Activated carbon/metal-organic framework composite as a bio-based novel green adsorbent: Preparation and mathematical pollutant removal modeling. J. Mol. Liq. 2019, 277, 310–322. [Google Scholar] [CrossRef]
  3. Singh, K.P.; Gupta, S.; Singh, A.K.; Sinha, S. Optimizing adsorption of crystal violet dye from water by magnetic nano composite using response surface modeling approach. J. Hazard. Mater. 2011, 186, 1462–1473. [Google Scholar] [CrossRef] [PubMed]
  4. Roosta, M.; Ghaedi, M.; Daneshfar, A.; Sahraei, R.; Asghari, A. Optimization of the ultrasonic assisted removal of methylene blue by gold nanoparticles loaded on activated carbon using experimental design methodology. Ultrason. Sonochemistry 2014, 21, 242–252. [Google Scholar] [CrossRef]
  5. Albadarin, A.; Collins, M.N.; Naushad, M.; Shirazian, S.; Walker, G.; Mangwandi, C. Activated lignin-chitosan extruded blends for efficient adsorption of methylene blue. Chem. Eng. J. 2017, 307, 264–272. [Google Scholar] [CrossRef] [Green Version]
  6. Marrakchi, F.; Khanday, W.A.; Asif, M.; Hameed, B.H. Cross-linked chitosan/sepiolite composite for the adsorption of methylene blue and reactive orange 16. Int. J. Biol. Macromol. 2016, 93, 1231–1239. [Google Scholar] [CrossRef]
  7. Viscusi, G.; Lamberti, E.; Gorrasi, G. Design of sodium alginate/soybean extract beads loaded with hemp hurd and halloysite as novel and sustainable systems for methylene blue adsorption. Polym. Eng. Sci. 2022, 62, 129–144. [Google Scholar] [CrossRef]
  8. Li, C.D.; Lu, J.J.; Li, S.M.; Tong, Y.B.; Ye, B.C. Synthesis of magnetic microspheres with sodium alginate and activated carbon for removal of methylene blue. Materials 2017, 10, 84. [Google Scholar] [CrossRef]
  9. Hassan, A.F.; Mohsen, A.M.; Fouda, M.M. Comparative study of calcium alginate, activated carbon, and their composite beads on methylene blue adsorption. Carbohyd. Polym. 2014, 102, 192–198. [Google Scholar] [CrossRef]
  10. Chen, T.; Zhao, Y.; Sang, Y.N.; Tang, M.; Hu, G.W.; Han, X.B.; Gao, J.; Ma, R. Facile synthesis of magnetic CS-g-SPSS microspheres via electron beam radiation for efficient removal of methylene blue. J. Saudi Chem. Soc. 2021, 25, 101299. [Google Scholar] [CrossRef]
  11. Hao, C.Q.; Li, G.F.; Wang, G.L.; Chen, W.; Wang, S.S. Preparation of acrylic acid modified alkalized MXene adsorbent and study on its dye adsorption performance. Colloids Surf. A Physicochem. Eng. Asp. 2022, 632, 127730. [Google Scholar] [CrossRef]
  12. Han, X.B.; Li, R.; Miao, P.P.; Gao, J.; Hu, G.W.; Zhao, Y.; Chen, T. Design, synthesis and adsorption evaluation of bio-based lignin/chitosan beads for congo red removal. Materials 2022, 15, 2310. [Google Scholar] [CrossRef]
  13. Hu, T.; Liu, Q.Z.; Gao, T.T.; Dong, K.J.; Wei, G.; Yao, J.S. Facile preparation of tannic acid-poly(vinyl alcohol)/sodium alginate hydrogel beads for methylene blue removal from simulated solution. ACS Omega 2018, 3, 7523–7531. [Google Scholar] [CrossRef]
  14. Kaczmarek, B. Improving sodium alginate films properties by phenolic acid addition. Materials 2020, 13, 2895. [Google Scholar] [CrossRef]
  15. Larosa, C.; Salerno, M.; Lima, J.S.; Meri, R.M.; Silva, M.F.; Carvalho, L.B.; Converti, A. Characterisation of bare and tannase-loaded calcium alginate beads by microscopic, thermogravimetric, FTIR and XRD analyses. Int. J. Biol. Macromol. 2018, 115, 900–906. [Google Scholar] [CrossRef]
  16. Abulateefeh, S.R.; Taha, M.O. Enhanced drug encapsulation and extended release profiles of calcium-alginate nanoparticles by using tannic acid as a bridging cross-linking agent. J. Microencapsul. 2015, 32, 96–105. [Google Scholar] [CrossRef]
  17. Ma, M.S.; Liu, Z.; Hui, L.F.; Shang, Z.; Yuan, S.Y.; Dai, L.; Liu, P.T.; Liu, X.L.; Ni, Y.H. Lignin-containing cellulose nanocrystals/sodium alginate beads as highly effective adsorbents for cationic organic dyes. Int. J. Biol. Macromol. 2019, 139, 640–646. [Google Scholar] [CrossRef]
  18. Wu, M.; Chen, W.J.; Mao, Q.H.; Bai, Y.S.; Ma, H.Z. Facile Synthesis of chitosan/gelatin filled with graphene bead adsorbent for orange II removal. Chem. Eng. Res. Des. 2019, 144, 35–46. [Google Scholar] [CrossRef]
  19. Liao, J.J.; Brosse, N.; Hoppe, S.; Du, G.B.; Zhou, X.J.; Pizzi, A. One-step compatibilization of poly(lactic acid) and tannin via reactive extrusion. Mater. Des. 2020, 191, 108603. [Google Scholar] [CrossRef]
  20. Zhu, X.Y.; Hou, X.L.; Ma, B.M.; Xu, H.L.; Yang, Q.Y. Chitosan/gallnut tannins composite fifiber with improved tensile, antibacterial and flfluorescence properties. Carbohydr. Polym. 2019, 226, 115311. [Google Scholar] [CrossRef]
  21. Ofomaja, A.E.; Ho, Y.S. Equilibrium sorption of anionic dye from aqueous solution by palm kernel fibre as sorbent. Dye. Pigment. 2007, 74, 60–66. [Google Scholar] [CrossRef]
  22. Zhang, L.J.; Zhao, D.Q.; Lu, Y.; Chen, J.H.; Li, H.T.; Xie, J.H.; Xu, Y.; Yuan, H.K.; Liu, X.J.; Zhu, X.Y.; et al. A Graphene oxide modified cellulose nanocrystal/PNIPAAm IPN hydrogel for the adsorption of congo red and methylene blue. New J. Chem. 2021, 45, 16679–16688. [Google Scholar] [CrossRef]
  23. Zhu, H.Y.; Fu, Y.Q.; Jiang, R.; Jiang, J.H.; Xiao, L.; Zeng, G.M.; Zhao, S.L.; Wang, Y. Adsorption removal of congo red onto magnetic cellulose/Fe3O4/activated carbon composite: Equilibrium, kinetic and thermodynamic studies. Chem. Eng. J. 2011, 173, 494–502. [Google Scholar] [CrossRef]
  24. Igwegbe, C.A.; Ighalo, J.O.; Onyechi, K.K.; Onukwuli, O.D. Adsorption of Congo red and malachite green using H3PO4 and NaCl-modifed activated carbon from rubber (Hevea brasiliensis) seed shells. Sustain. Water Resour. Manag. 2021, 7, 63. [Google Scholar] [CrossRef]
  25. Igwegbe, C.A.; Ighalo, J.O.; Ghosh, S.; Ahmadi, S.; Ugonabo, V.I. Pistachio (Pistacia vera) waste as adsorbent for wastewater treatment: A review. Biomass Convers. Biorefinery 2021. [Google Scholar] [CrossRef]
  26. Luo, J.M.; Fu, K.X.; Yu, D.Y.; Hristovski, K.D.; Westerhoff, P.; Crittenden, J.C. Review of advances in engineering nanomaterial adsorbents for metal removal and recovery from water: Synthesis and microstructure impacts. ACS EST Eng. 2021, 1, 623–661. [Google Scholar] [CrossRef]
  27. Zhang, C.L.; Chen, Z.Z.; Guo, W.; Zhu, C.W.; Zou, Y.J. Simple fabrication of chitosan/graphene nanoplates composite spheres for efficient adsorption of acid dyes from aqueous solution. Int. J. Biol. Macromol. 2018, 112, 1048–1054. [Google Scholar] [CrossRef]
  28. Gu, F.; Geng, J.; Li, M.L.; Chang, J.M.; Cui, Y. Synthesis of chitosan-ignosulfonate composite as an adsorbent for dyes and metal ions removal from wastewater. ACS Omega 2019, 4, 21421–21430. [Google Scholar] [CrossRef] [Green Version]
  29. Duan, Y.Q.; Freyburger, A.; Kunz, W.; Zollfrank, C. Lignin/chitin films and their adsorption characteristics for heavy metal ions. ACS Sustain. Chem. Eng. 2018, 6, 6965–6973. [Google Scholar] [CrossRef]
  30. Sartape, A.S.; Mandhare, A.M.; Salvi, P.P.; Pawar, D.K.; Kolekar, S.S. Kinetic and equilibrium studies of the adsorption of Cd(II) from aqueous solutions by wood apple shell activated carbon. Desalination Water Treat. 2013, 51, 4638–4650. [Google Scholar] [CrossRef]
  31. Chen, T.; Yan, C.J.; Wang, Y.X.; Tang, C.H.; Zhou, S.; Zhao, Y.; Ma, R.; Duan, P. Synthesis of activated carbon-based amino phosphonic acid chelating resin and its adsorption properties for Ce(III) removal. Environ. Technol. 2015, 36, 2168–2176. [Google Scholar] [CrossRef] [PubMed]
  32. Wu, F.C.; Tseng, R.L.; Juang, R.S. Characteristics of Elovich equation used for the analysis of adsorption kinetics in dye-chitosan systems. Chem. Eng. J. 2009, 150, 366–373. [Google Scholar] [CrossRef]
  33. Yeddou, N.; Bensmaili, A. Kinetic models for the sorption of dye from aqueous solution by clay-wood sawdust mixture. Desalination 2005, 185, 499–508. [Google Scholar] [CrossRef]
  34. Aljar, M.A.A.; Rashdan, S.; Fattah, A.A. Environmentally friendly polyvinyl alcohol-alginate/bentonite semi-interpenetrating polymer network nanocomposite hydrogel beads as an effificient adsorbent for the removal of methylene blue from aqueous solution. Polymers 2021, 13, 4000. [Google Scholar] [CrossRef]
  35. Yu, D.Y.; Wang, Y.J.; Wu, M.H.; Zhang, L.; Wang, L.L.; Ni, H.G. Surface functionalization of cellulose with hyperbranched polyamide for effificient adsorption of organic dyes and heavy metals. J. Clean. Prod. 2019, 232, 774–783. [Google Scholar] [CrossRef]
Figure 1. Bio-based Tan/SA hydrogel absorbent for the removal of MB.
Figure 1. Bio-based Tan/SA hydrogel absorbent for the removal of MB.
Gels 08 00486 g001
Figure 2. Photos of SA beads (a) and Tan/SA-40% beads (d), SEM images of SA beads (b,c), Tan/SA-40% beads (e,f).
Figure 2. Photos of SA beads (a) and Tan/SA-40% beads (d), SEM images of SA beads (b,c), Tan/SA-40% beads (e,f).
Gels 08 00486 g002
Figure 3. FTIR spectra of SA, Tan, 40% Tan/SA beads.
Figure 3. FTIR spectra of SA, Tan, 40% Tan/SA beads.
Gels 08 00486 g003
Figure 4. SEM images of SA hydrogel beads (a,b), Tan/SA-40% hydrogel beads (c,d) after heat treatment (T = 50 °C, t = 48 h, s = 200 rpm).
Figure 4. SEM images of SA hydrogel beads (a,b), Tan/SA-40% hydrogel beads (c,d) after heat treatment (T = 50 °C, t = 48 h, s = 200 rpm).
Gels 08 00486 g004
Figure 5. TG spectra of SA, Tan and 40% Tan/SA beads.
Figure 5. TG spectra of SA, Tan and 40% Tan/SA beads.
Gels 08 00486 g005
Figure 6. Effect of the initial pH on the removal efficiency with 40% Tan/SA beads (C0 = 80 mg/L, T = 25 °C, t = 4 h).
Figure 6. Effect of the initial pH on the removal efficiency with 40% Tan/SA beads (C0 = 80 mg/L, T = 25 °C, t = 4 h).
Gels 08 00486 g006
Figure 7. Effect of the temperature on the removal efficiency with 40% Tan/SA beads (C0 = 80 mg/L, pH = 7, t = 4 h).
Figure 7. Effect of the temperature on the removal efficiency with 40% Tan/SA beads (C0 = 80 mg/L, pH = 7, t = 4 h).
Gels 08 00486 g007
Figure 8. Effect of Tan content on the adsorption capacity (C0 = 200 mg/L, pH = 12, T = 45 °C, t = 24 h).
Figure 8. Effect of Tan content on the adsorption capacity (C0 = 200 mg/L, pH = 12, T = 45 °C, t = 24 h).
Gels 08 00486 g008
Figure 9. Effect of initial concentration and contact time on the adsorption capacity Q (mg/g) with 40%Tan/SA beads (pH = 12, T = 45 °C).
Figure 9. Effect of initial concentration and contact time on the adsorption capacity Q (mg/g) with 40%Tan/SA beads (pH = 12, T = 45 °C).
Gels 08 00486 g009
Figure 10. Fitting of pseudo-first-order (a), pseudo-second-order (b), and Elovich (c) kinetic models for the MB adsorption.
Figure 10. Fitting of pseudo-first-order (a), pseudo-second-order (b), and Elovich (c) kinetic models for the MB adsorption.
Gels 08 00486 g010
Figure 11. Effect of recycling times on the adsorption capacity Q (mg/g) with 40%Tan/SA beads (C0. = 10 mg/L, pH = 12, T = 45 °C, t = 4 h).
Figure 11. Effect of recycling times on the adsorption capacity Q (mg/g) with 40%Tan/SA beads (C0. = 10 mg/L, pH = 12, T = 45 °C, t = 4 h).
Gels 08 00486 g011
Table 1. Thermodynamic parameters for the adsorption of MB on 40% Tan/SA beads.
Table 1. Thermodynamic parameters for the adsorption of MB on 40% Tan/SA beads.
Temperature (K)K0ΔG0 (kJ/mol)ΔH0 (kJ/mol)ΔS0 (kJ/mol·k)
2880.362.4480.800.27
2980.690.92--
3082.53−2.38--
3188.25−5.58--
Table 2. Kinetic parameters of three models for MB adsorption.
Table 2. Kinetic parameters of three models for MB adsorption.
Kinetic ModelsCoefficients60 mg/L80 mg/L100 mg/L
Pseudo-first-orderqe,cal (mg/g)9.727915.879027.4949
k1 (min1)0.01070.01390.0172
R20.90290.80500.9666
Pseudo-second-orderqe,cal (mg/g)86.81116.55144.93
k2 (×10−3) (g/mg min)2.782.071.31
R20.99990.99990.9999
Elovich modelα (mg/g min)3.14 × 1065.87 × 1062.89 × 105
β (g/mg)0.2190.1660.110
R20.81470.73010.8463
Table 3. Comparison of the MB removal performance with similar adsorbents.
Table 3. Comparison of the MB removal performance with similar adsorbents.
AdsorbentC0
(mg/L)
Qmax
(mg/g)
Removal
(%)
RegenerationRef.
Tan/SA200247.282.4>90%, 5 cyclesThis work
Lig/CS8236.2588.4ND[5]
CS/SP30040.98627.3ND[6]
SA/SB/HNT1004994ND[7]
Fe3O4/AC/SA700222.331.8ND[8]
SA/AC100073073ND[9]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gao, J.; Li, Z.; Wang, Z.; Chen, T.; Hu, G.; Zhao, Y.; Han, X. Facile Synthesis of Sustainable Tannin/Sodium Alginate Composite Hydrogel Beads for Efficient Removal of Methylene Blue. Gels 2022, 8, 486. https://doi.org/10.3390/gels8080486

AMA Style

Gao J, Li Z, Wang Z, Chen T, Hu G, Zhao Y, Han X. Facile Synthesis of Sustainable Tannin/Sodium Alginate Composite Hydrogel Beads for Efficient Removal of Methylene Blue. Gels. 2022; 8(8):486. https://doi.org/10.3390/gels8080486

Chicago/Turabian Style

Gao, Jie, Zhenzhen Li, Ziwen Wang, Tao Chen, Guowen Hu, Yuan Zhao, and Xiaobing Han. 2022. "Facile Synthesis of Sustainable Tannin/Sodium Alginate Composite Hydrogel Beads for Efficient Removal of Methylene Blue" Gels 8, no. 8: 486. https://doi.org/10.3390/gels8080486

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop