Home Adsorption of l-α-glycerophosphocholine on ion-exchange resin: Equilibrium, kinetic, and thermodynamic studies
Article Open Access

Adsorption of l-α-glycerophosphocholine on ion-exchange resin: Equilibrium, kinetic, and thermodynamic studies

  • Hongya Li EMAIL logo , Biao Yan , Yajun Ma , Xiangrong Ma , Xiaoli Zhang and Binxia Zhao
Published/Copyright: June 2, 2020
Become an author with De Gruyter Brill

Abstract

The adsorption of l-α-glycerophosphocholine (GPC) by cation-exchange resin 001 × 7 was studied in a batch system. The adsorbent dosage, shaking speed, and adsorption temperature were investigated. An adsorption efficiency of more than 99.4% was obtained under optimal conditions. The kinetic data evaluated by the pseudo-second-order kinetic model fitted the experimental data better than those evaluated by the pseudo-first-order model. The rate constant k2 increased when the temperature increased, indicating the adsorption was endothermic in nature. The Langmuir and Freundlich isotherm models were used to analyze the adsorption equilibrium data, and it was found that the experimental data well fitted the Langmuir isotherm model. The thermodynamic parameters, enthalpy change (ΔG0), free energy change (ΔH0), and entropy change (ΔS0), were calculated. The value of ΔG0 was found to be in the range of −5.09 to −14.20 kJ mol−1, indicating that the adsorption was spontaneous and basically physisorption, and the positive values of ΔH0 and ΔS0 exhibited that the adsorption was endothermic and the randomness of the system increased during the adsorption.

1 Introduction

Ion-exchange resin is a spherical polymer compound with different functional groups [1], which are connected with the insoluble three-dimensional spatial network backbone. It has received significant attention because of its high adsorption speed, high mechanical strength, better stability, and reutilization. In the past few years, ion-exchange resins have been studied extensively and have been proven to be most effective in the process of separation, purification, and other biomedical applications [1,2]. Zhang et al. [3] selected D101 from nine different kinds of macroporous resins to purify flavonoids, and results showed that the recovery of flavonoids can be above 93% under optimum conditions. Shi et al. [4] proved that three weakly basic anion-exchange resins: D301, D314, and D354, can be efficiently used for the removal of hexavalent chromium from aqueous solutions. Ion-exchange resins can be widely used in water treatment, environmental protection, food processing, pharmaceutical, petrochemical, and other fields.

l-α-Glycerophosphocholine (GPC, for short) is widespread in all our cells, and it is important for our mental focus, attention, and other higher mental functions, so it is widely used in cosmetics, pharmaceutical field, and other applications [5,6,7]. Compared with the biological extraction method and the chemical synthesis method, hydrolysis of natural lecithin is the perfect method for preparing GPC, because, due to the natural origin, the product is more credible [8,9,10]. However, many studies used the solution of tetrabutylammonium hydroxide [11] or sodium methoxide [12] as a catalyst, and the catalyst is difficult to separate, so the purification of GPC is the most complicated step for its pharmaceutical application.

GPC contains many hydroxyl groups and can easily dissolve in aqueous solutions or alcohols. It can also be purified by an ion-exchange resin [1,13]. Resin 001 × 7 is one of the strong-acidic cation-exchange resins, and its adsorption capacity is equivalent to that of Amberlite IR-120 [1]. In this study, the 001 × 7 resin was chosen as a potential adsorbent for adsorbing GPC. Factors influencing the adsorption effect, adsorbent dosage, shaking speed, GPC concentration, and adsorption temperature, were evaluated thoroughly. The adsorption kinetics was determined using pseudo-first-order and pseudo-second-order models, and the equilibrium isotherm was fitted by Langmuir and Freundlich adsorption isotherm models. This preliminary study can provide theoretical basis and technical support for further separation investigations [1].

2 Materials and methods

2.1 Materials

GPC was purchased from Shanghai Yi Yao Fine Chemical Plant (Shanghai, P. R. China) and stored under airtight and light-free conditions until use. The resin 001 × 7 was obtained from Sunresin New Materials Co., Ltd (Xi’an, P. R. China); it was treated according to the manufacturer’s instructions and then washed with anhydrous methanol several times to remove the residual water in it, and finally, the resin was kept in anhydrous methanol at normal temperature. Hydrochloric acid, ammonia, n-propanol, methanol, and chloroform were purchased from Sinopharm Chemical Reagent Co., Ltd (Shanghai, P. R. China). All chemicals used were of analytical grade.

2.2 Adsorption experiments

The adsorption experiments were performed in stoppered conical flasks in a water bath shaker, the cation-exchange resin 001 × 7 was chosen as a potential adsorbent, and solution of GPC was prepared in anhydrous methanol. The experiments were conducted with different resin amounts (wet-based weight) between 0.15 and 0.30 g mL−1 to optimize the solid/liquid ratio [14], and the effect of shaking speed on the adsorption process was investigated in the range of 120–240 rpm. The temperature was investigated from 20 to 40°C. During the experiments, the samples were drawn from the solution at predetermined time intervals before the adsorption equilibrium was achieved. All experiments were carried out in triplicate to minimize the random error.

2.3 Analytical method

High-performance liquid chromatography (HPLC) and thin-layer chromatography (TLC) determination methods were established and compared. The results showed that the minimum detection limit of HPLC was a bit lower than that of TLC; that is, the HPLC method is slightly more sensitive than the TLC method. However, the scanned data of TLC within 10 h after spotting were similar to those of HPLC; at the same time, the TLC is a simple, fast, and low-cost method, and it can detect more samples at the same time, so the TLC method is used in the experiments. The samples were treated by centrifugation at 10,000 rpm for 5 min, then a certain amount of each sample (including the GPC solution before adsorption) was spotted on the TLC plate and developed in the solvent system of n-propanol:ammonia (5 mol L−1) = 13:7. The developed plate was detected in an iodine chamber after drying. Some yellow spots appeared on the white background after a period of time. Finally, the spots were quantified by densitometric scanning with a Shimadzu Dual-Wavelength TLC Scanner CS-930. The amount of GPC adsorbed, qe, was computed by the following expression:

(1)qe=c0cemV,

where c0 and ce (mg L−1) are the concentration of GPC solution before and after adsorption, V is the volume of GPC solution (L), and m is the mass of wet cation-exchange resin (g). The adsorption efficiency of GPC was calculated by the following equation [15]:

(2)Adsorptionefficiency%=c0cec0×100%.

3 Results and discussion

3.1 Effect of adsorbent dosage

The effect of adsorbent dosage on adsorption of GPC on the resin 001 × 7 was studied by varying the amount of resin in the medium from 0.15 to 0.30 g mL−1 under the following conditions: GPC initial concentration, 7,450 mg L−1; adsorption temperature, 25°C; shaking speed, 120 rpm; and contact time, 2 h. The results are presented in Figure 1. The adsorption efficiency for GPC increased proportionally with an increase in the amount of resin. It increased from 82% to 98% when the adsorbent dosage increased from 0.15 to 0.20 g mL−1, because when the adsorbent dosage increased, the surface area of the resin and its available adsorption sites for GPC molecules increased, consequently resulting in better adsorption [16,17]. But no obvious change was observed above the adsorbent dosage of 0.20 g mL−1, so in the subsequent experiments, the appropriate adsorbent dosage was 0.20 g mL−1.

Figure 1 Effect of resin dosage on the adsorption efficiency for GPC by resin 001 × 7.
Figure 1

Effect of resin dosage on the adsorption efficiency for GPC by resin 001 × 7.

3.2 Effect of shaking speed

The shaking speed is also one of the most important parameters which would affect the distribution of the adsorbate ions. The dependence of GPC adsorption on shaking speed was studied by varying the shaking speed (120, 160, 200, and 240 rpm), while the other factors such as the initial GPC concentration, adsorption temperature, and contact time remained constant. The results are shown in Figure 2. It was apparent that when the shaking speed increased from 120 to 240 rpm, the adsorption efficiency for GPC increased from 42% to 58% in the initial 10 min. This is because the uniformity of the system increased and the film resistance to mass transfer decreased [18]. But the final results did not change significantly in the experimental shaking speed. As clarified by Dotto and Pinto [19], there was usually a slight change in adsorption behavior when the shaking speed was between 100 and 200 rpm.

Figure 2 Effect of shaking speed on the adsorption efficiency for GPC by 001 × 7 cation-exchange resin.
Figure 2

Effect of shaking speed on the adsorption efficiency for GPC by 001 × 7 cation-exchange resin.

3.3 Effect of adsorption temperature

To determine the dependence of GPC adsorption on temperature, experiments were performed in the temperature range of 20 to 40°C. A shaking speed of 120 rpm was selected, and the other parameters remained constant. The results are shown in Figure 3.

Figure 3 Effect of temperature on the adsorption efficiency for GPC by 001 × 7 cation-exchange resin.
Figure 3

Effect of temperature on the adsorption efficiency for GPC by 001 × 7 cation-exchange resin.

It could be found that the adsorption process reached equilibrium state quickly (30 min) when the temperature was above 30°C. The amount of GPC adsorbed on the resin increased when the adsorption temperature changed from 20 to 40°C, indicating that the adsorption of GPC on the resin 001 × 7 was favored at higher temperature. This may be a result that the higher temperature increased the mobility of GPC molecules which can provide an extra driving force for GPC molecules to transfer from the aqueous to solid phases [20,21].

3.4 Adsorption kinetic studies

In order to examine the controlling mechanism of adsorption and to control the process efficiently, the kinetic studies were very necessary. The adsorption kinetics could provide important information for studying the mechanism [22] and selecting optimum operation conditions [23]. The kinetics of GPC adsorption on the resin 001 × 7 was examined with two familiar kinetic models: the pseudo-first-order and pseudo-second-order reaction kinetic models. The pseudo-first-order reaction kinetic model of Lagergren was widely used, and the simple linear equation is given below [24,25]:

(3)ln(qeqt)=lnqek1t,

where qe and qt are the amounts of GPC adsorbed at equilibrium and at time t (mg g−1) and k1 is the pseudo-first-order rate constant (min−1).

Plots of ln(qeqt) against t were used to determine the rate constants k1 and correlation coefficients R2 for different adsorbent dosages and temperatures (Figure 4).

Figure 4 Pseudo-first-order plots of GPC adsorption on resin 001 × 7 at different (a) adsorbent dosages and (b) temperatures.
Figure 4

Pseudo-first-order plots of GPC adsorption on resin 001 × 7 at different (a) adsorbent dosages and (b) temperatures.

In addition, the pseudo-second-order kinetic equation is expressed as [24,25]:

(4)tqt=1k2qe2+tqe,

where k2 (g mg−1 min−1) is the pseudo-second-order rate constant.

Similar to the pseudo-first-order reaction kinetic model, qe and the rate constant k2 can be obtained from the slope and intercepts of plots t/qt versus t (Figure 5).

Figure 5 Pseudo-second-order plots of GPC adsorption on resin 001 × 7 at different (a) adsorbent dosages and (b) temperatures.
Figure 5

Pseudo-second-order plots of GPC adsorption on resin 001 × 7 at different (a) adsorbent dosages and (b) temperatures.

The calculated adsorbed amounts of GPC at equilibrium (qe,cal), rate constant (k1, k2), and the correlation coefficient (R2) for both pseudo-first-order and pseudo-second-order reaction kinetic models are listed in Table 1. The values of qe,cal from the pseudo-first-order reaction kinetic model were significantly different from the experimental values (qe,exp), and the correlation coefficients (R2), in the range of 0.6081–0.9773, were also lower, implying that the pseudo-first-order kinetic model was not suitable for this sorption system. While for the pseudo-second-order equation, the qe,cal values agreed well with the experimental values qe,exp, and the correlation coefficients were all greater than 0.99 at different adsorbent dosages and temperatures.

Table 1

Kinetic parameters for the adsorption of GPC on resin 001 × 7 based on the pseudo-first-order and pseudo-second-order kinetic equations at different adsorbent dosages and temperatures

Adsorbent dosage (g mL−1)Temperature (°C)qe,exp (mg g−1)Pseudo-first-orderPseudo-second-order
qe,cal (mg g−1)k1 (min−1)R2qe,cal (mg g−1)k2 (g mg−1 min−1)R2
0.1540.8263.780.06800.925943.860.00280.9931
0.2036.8644.160.06320.865038.460.00490.9976
0.2529.8319.290.06350.977330.860.00930.9983
0.3024.8610.660.07400.861225.130.03540.9998
2023.318.870.04160.658727.170.00710.9903
2528.6520.010.05570.870730.210.00770.9906
3029.8210.890.05680.826129.940.02520.9991
3529.836.250.06000.731829.850.07190.9999
4029.833.010.05980.608129.760.29710.9999

On the other hand, the plots in Figure 5 show better linearity than those in Figure 4. Both facts suggest that the adsorption of GPC by the resin 001 × 7 followed the pseudo-second-order kinetic model. The values of k2 increased from 0.0071 to 0.2971 g mg−1 min−1 as the temperature increased from 20 to 40°C, showing that the GPC adsorption on the resin 001 × 7 was an endothermic process.

3.5 Adsorption isotherms

In general, the adsorption isotherms are important to estimate the mechanism of the process [26,27]. The experimental data for GPC adsorption isotherms at different temperatures are shown in Figure 6. The adsorption capacity for GPC increased when the initial concentration increased; however, the higher adsorption efficiency was obtained at the lower concentration. This can be because the number of active sites on the per-unit surface area of resin reduced at higher GPC concentration.

Figure 6 Plots of ce against qe for the GPC adsorption on resin 001 × 7 at different temperatures.
Figure 6

Plots of ce against qe for the GPC adsorption on resin 001 × 7 at different temperatures.

Analysis of equilibrium data is important to compare different sorbents and to optimize an operating procedure [28,29]. The equilibrium data obtained in the experiments were analyzed with well-known adsorption models such as Langmuir and Freundlich models [18,30]. The Langmuir isotherm model (equation (5)) supposes that the adsorption arises on a homogeneous surface and there is no lateral interaction between the adsorbed molecules. The Freundlich isotherm (equation (6)) is usually used to analyze the adsorption on a heterogeneous surface. The linear form of them is expressed by the following equations:

(5)ceqe=1qmaxb+ceqmax
(6)lnqe=lnKf+1nlnce,

where ce is the concentration of GPC at equilibrium in solution (mg L−1); qmax is the maximum adsorption capacity of adsorbent (mg g−1); b is the adsorption energy coefficient (L mg−1); and Kf (mg g−1) and n are the Freundlich constants, respectively.

Linear plots of the two models of GPC adsorption on the resin 001 × 7 are shown in Figures 7 and 8. The plots in Figure 7 show better linearity than those in Figure 8. The isotherm constants for adsorption of GPC on the resin 001 × 7 at different temperatures are shown in Table 2. The detailed analysis of the correlation coefficient R2 values showed that the Langmuir model (0.9993–0.9999) fit the experimental data better than the Freundlich model (0.8233–0.9230) at different temperatures. From Table 2, it is also clear that the qmax increased when the temperature increased, once again indicating the adsorption was an endothermic process.

Figure 7 Langmuir isotherm plots for the adsorption of GPC on resin 001 × 7 at different temperatures.
Figure 7

Langmuir isotherm plots for the adsorption of GPC on resin 001 × 7 at different temperatures.

Figure 8 Freundlich isotherm plots for the adsorption of GPC on resin 001 × 7 at different temperatures.
Figure 8

Freundlich isotherm plots for the adsorption of GPC on resin 001 × 7 at different temperatures.

Table 2

Adsorption isotherm constants for GPC adsorption on resin 001 × 7

Temperature (°C)LangmuirFreundlich
qmax (mg g−1)b (L mg−1)R2RLaKFnR2
2029.150.01290.99930.014815.1513.250.8233
2531.850.07600.99990.002619.7616.980.8278
3037.880.03790.99990.005118.2110.920.9230
3543.290.03420.99990.005717.298.440.9011
  1. a

    Initial concentration (C0): 5,144 mg L−1.

The essential features of the Langmuir isotherm are expressed by a dimensionless constant separation factor RL, and it can be expressed as follows [31,32]:

(7)RL=11+bc0,

where c0 is the initial concentration of GPC (mg L−1) and b is the Langmuir adsorption energy constant (L mg−1). The value of RL indicates the shape of isotherm to be either irreversible (RL = 0) or favorable (0 < RL < 1) or linear (RL = 1) or unfavorable (RL > 1) [33,34,35]. The results showed that the RL values (listed in Table 2) were all in the range of 0–1 at 20–35°C, indicating that the cation-exchange resin 001 × 7 is favorable for adsorption of GPC under the studied conditions [33].

3.6 Thermodynamic parameters

In order to evaluate the adsorption process better, the thermodynamic parameters, standard free energy change ΔG0, standard enthalpy change ΔH0, and standard entropy change ΔS0, were also obtained. They were calculated by using the following equations [36]:

(8)Kc=cAecSe
(9)ΔG0=RTlnKc
(10)lnKc=ΔS0RΔH0RT,

where Kc is the equilibrium constant, and cSe and cAe are the equilibrium concentration of GPC on resin and in solution (mg L−1), respectively.

The values of ΔH0 and ΔS0 can be obtained from the slope and intercept of the plot of ln Kc versus 1/T (Figure 9). All the thermodynamic parameters were calculated and are tabulated in Table 3. The values of ΔG0 were negative at all experiment temperatures, indicating that the adsorption process was spontaneous. According to Crini and Badot [37], the more negative value of ΔG0 at higher temperature implies a greater driving force for adsorption; therefore, the adsorption is more favorable at higher temperature. The ΔG0 for physical adsorption is between −20 and 0 kJ mol−1 usually, and for chemical adsorption, it is between −400 and −80 kJ mol−1 [36]. The ΔG0 for GPC adsorption was between −5.09 and −14.20 kJ mol−1, so it was basically physisorption. The value of ΔH0 was positive, exhibiting that the adsorption was endothermic, and it was anticipated that the uptake of GPC from solution increased when the solution temperature increased. The positive values of ΔS0 indicated that the randomness at the interface of resin and solution increased during the adsorption.

Figure 9 Variation of equilibrium constant (Kc) as a function of temperature.
Figure 9

Variation of equilibrium constant (Kc) as a function of temperature.

Table 3

Thermodynamic parameters for the adsorption of GPC on resin 001 × 7

Temperature (°C)ΔG0 (kJ mol−1)ΔH0 (kJ mol−1)ΔS0 (J K−1 mol−1)
20−5.09129.70460.01
25−7.89
30−10.57
35−12.62
40−14.20

4 Conclusion

This preliminary study showed that the resin 001 × 7 can be used as a potential adsorbent for adsorbing GPC. The optimum conditions for maximum adsorption of GPC were found to be as follows: adsorbent dosage, 0.20 g mL−1; GPC initial concentration, 7,450 mg L−1; adsorption temperature, 25°C; and shaking speed, 160 rpm. The kinetics of GPC adsorption obeyed the pseudo-second-order kinetic model, and the adsorption was an endothermic process because the rate constant k2 increased when the temperature increased. The adsorption data fitted the Langmuir model better than the Freundlich model. The values of ΔG0, ΔH0, and ΔS0 were calculated to predict the kinetic aspects of GPC adsorption. The negative values of ΔG0 at the experimental temperatures indicated that the adsorption was spontaneous, and the positive values of ΔH0 and ΔS0 exhibited that the adsorption was endothermic and the randomness increased during the adsorption.

Acknowledgments

This study was funded by the National Natural Science Foundation of China (grant no. 21663033 and 21763030), the Yulin City Industry University Research Cooperation Project (grant no. 2016CXY-05-01), and the Startup Foundation for Advanced Talents of Yulin University (award no. 16GK10).

References

[1] Zhang JB, Zhang XL, Li HY, Zhao BX. Ion exchange adsorption studies of miglitol in a fixed bed. Chem Eng Technol. 2012;35(5):811–8.10.1002/ceat.201100685Search in Google Scholar

[2] Anand V, Kandarapu R, Garg S. Ion-exchange resins: carrying drug delivery forward. Drug Discov Today. 2001;6(17):905–14.10.1016/S1359-6446(01)01922-5Search in Google Scholar

[3] Zhang Y, Li SF, Wu XW, Zhao X. Macroporous resin adsorption for purification of flavonoids in houttuynia cordata thunb. Chinese J Chem Eng. 2007;15(6):872–6.10.1016/S1004-9541(08)60017-8Search in Google Scholar

[4] Shi TH, Wang ZC, Liu Y, Jia SG, Du CM. Removal of hexavalent chromium from aqueous solutions by D301, D314 and D354 anion-exchange resins. J Hazard Mater. 2009;161(2–3):900–6.10.1016/j.jhazmat.2008.04.041Search in Google Scholar PubMed

[5] Ichihara K, Iwasaki H, Ueda K, Takizawa R, Naito H, Tomosugi M. Synthesis of phosphatidylcholine: an improved method without using the cadmium chloride complex of sn-glycero-3-phosphocholine. Chem Phys Lipids. 2005;137(1–2):94–9.10.1016/j.chemphyslip.2005.06.001Search in Google Scholar PubMed

[6] Li B, Li HY, Zhang XL, Fan PJ, Liu L, Li BL, et al. Calcined sodium silicate as an efficient and benign heterogeneous catalyst for the transesterification of natural lecithin to l-α-glycerophosphocho-line. Green Process Synth. 2019;8:78–84.10.1515/gps-2017-0190Search in Google Scholar

[7] Zhang TT, Zhang XL, Li HY, Bai WL, Zhao BX. Kinetic evaluation of the ethanolysis of phosphatidylcholine catalyzed by choline hydroxide. React Kinet Mech Cat. 2013;110(1):31–9.10.1007/s11144-013-0579-9Search in Google Scholar

[8] Li HY, Zhang XL, Yan B, Zhao BX, Gao J. Preparation and catalytic performance of quaternary ammonium base resin for methanolysis of natural phosphatidylcholine. Chem Pap. 2016;70(6):706–12.10.1515/chempap-2016-0023Search in Google Scholar

[9] Song YS, Song ES, Kang DS, Song IW, Kang PG, Oh SS, et al. A process for preparation of l-alpha-glycerophosphoryl choline. WO Pat., 2007/145476A1; 2007.Search in Google Scholar

[10] Li HY, Zhang XL, Zhang JB, Zhang TT, Zhao BX. Preparing l-α-glycerophosphocholine in a phase-transfer catalytic reaction: kinetic study. Int J Chem Kinet. 2014;46(3):169–75.10.1002/kin.20840Search in Google Scholar

[11] Tronconi G, Coccoli C. Process for separating the main components of a mixture of raw deacylated phospholipids. WO Pat., 91/15494; 1991.Search in Google Scholar

[12] Tronconi G. Process for the preparation of l-α-glycerylphosphorylcholine and of l-α-glycerylphosphorylethanolamine. WO Pat., 90/13552; 1990.Search in Google Scholar

[13] Zhang KY, Wang XG, Huang JH, Liu YF. Purification of l-alpha glycerylphosphorylcholine by column chromatography. J Chromatogr A. 2012;1220:108–14.10.1016/j.chroma.2011.12.003Search in Google Scholar PubMed

[14] Bayramoglu G, Bektas S, Arica MY. Adsorption kinetics and thermodynamic parameters of cationic dyes from aqueous solutions by using a new strong cation-exchange resin. Chem Eng J. 2009;152(2–3):339–46.10.1016/j.cej.2009.04.051Search in Google Scholar

[15] Balarak D, Azarpira H, Mostafapour FK. Study of the adsorption mechanisms of cephalexin on to azolla filiculoides. Der Pharma Chemica. 2016;8(10):114–21.Search in Google Scholar

[16] Mahvi AH, Mostafapour FK. Biosorption of tetracycline from aqueous solution by azolla filiculoides: equilibrium kinetic and thermodynamics studies. Fresen Environ Bull. 2018;27(8):5759–67.Search in Google Scholar

[17] Balarak D, Mostafapour F, Bazrafshan E, Saleh TA. Studies on the adsorption of amoxicillin on multi-wall carbon nanotubes. Water Sci Technol. 2017;75(7):1599–606.10.2166/wst.2017.025Search in Google Scholar PubMed

[18] Albadarin AB, Mangwandi C, Al-Muhtaseb AH, Walker GM, Allen SJ, Ahmad MNM. Kinetic and thermodynamics of chromium ions adsorption onto low-cost dolomite adsorbent. Chem Eng J. 2012;179:193–202.10.1016/j.cej.2011.10.080Search in Google Scholar

[19] Dotto GL, Pinto LAA. Adsorption of food dyes acid blue 9 and food yellow 3 onto chitosan: stirring rate effect in kinetics and mechanism. J Hazard Mater. 2011;187(1–3):164–70.10.1016/j.jhazmat.2011.01.016Search in Google Scholar PubMed

[20] Balarak D, Azarpira H. Rice husk as a biosorbent for antibiotic metronidazole removal: isotherm studies and model validation. Int J ChemTech Res. 2016;9(7):566–73.Search in Google Scholar

[21] Azarpira H, Mahdavi Y, Khaleghi O. Thermodynamic studies on the removal of metronidazole antibiotic by multi-walled carbon nanotubes. Der Pharm Lett. 2016;8(11):107–13.Search in Google Scholar

[22] Ngah WSW, Fatinathan S. Adsorption characterization of Pb(ii) and Cu(ii) ions onto chitosan-tripolyphosphate beads: kinetic, equilibrium and thermodynamic studies. J Environ Manage. 2010;91(4):958–69.10.1016/j.jenvman.2009.12.003Search in Google Scholar PubMed

[23] Gupta VK, Sharma S. Removal of zinc from aqueous solutions using bagasse fly ash-a low cost adsorbent. Ind Eng Chem Res. 2003;42(25):6619–24.10.1021/ie0303146Search in Google Scholar

[24] Toor M, Bo J. Adsorption characteristics, isotherm, kinetics, and diffusion of modified natural bentonite for removing diazo dye. Chem Eng J. 2012;187:79–88.10.1016/j.cej.2012.01.089Search in Google Scholar

[25] Shankar P, Gomathi T, Vijayalakshmi K, Sudha PN. Comparative studies on the removal of heavy metals ions onto cross linked chitosan-g-acrylonitrile copolymer. Int J Biol Macromol. 2014;67:180–8.10.1016/j.ijbiomac.2014.03.010Search in Google Scholar PubMed

[26] Ahmadi S, Banach A, Kord Mostafapour F. Study survey of cupric oxide nanoparticles in removal efficiency of ciprofloxacin antibiotic from aqueous solution: adsorption isotherm study. Desalin Water Treat. 2017;89:297–303.10.5004/dwt.2017.21362Search in Google Scholar

[27] Balarak D, Azarpira H, Mostafapour FK. Adsorption isotherm studies of tetracycline antibiotics from aqueous solutions by maize stalks as a cheap biosorbent. Int J Pharm Technol. 2016;8(3):16664–75.Search in Google Scholar

[28] Bayramoglu G, Bektas S, Arica MY. Removal of Cd(ii), Hg(ii) and Pb(ii) ions from aqueous solution using P(HEMA/chitosan) membranes. J Appl Polym Sci. 2007;106(1):169–77.10.1002/app.26704Search in Google Scholar

[29] Roosta M, Ghaedi M, Daneshfar A, Sahraei R. Experimental design based response surface methodology optimization of ultrasonic assisted adsorption of safaranin o by tin sulfide nanoparticle loaded on activated carbon. Spectrochim Acta A. 2014;122:223–31.10.1016/j.saa.2013.10.116Search in Google Scholar PubMed

[30] Vimonses V, Lei S, Jina B, Chow CWK, Saint C. Kinetic study and equilibrium isotherm analysis of congo red adsorption by clay materials. Chem Eng J. 2009;148(2–3):354–64.10.1016/j.cej.2008.09.009Search in Google Scholar

[31] Hall KR, Eagleton LC, Acrivos A, Vermeulen T. Pore- and solid-diffusion kinetics in fixed-bed adsorption under constant-pattern conditions. Ind Eng Chem Fundamen. 1966;5(2):212–23.10.1021/i160018a011Search in Google Scholar

[32] Anirudhan TS, Radhakrishnan PG. Thermodynamics and kinetics of adsorption of Cu(ii) from aqueous solutions onto a new cation exchanger derived from tamarind fruit shell. J Chem Thermodyn. 2008;40(4):702–9.10.1016/j.jct.2007.10.005Search in Google Scholar

[33] Septhum C, Rattanaphani S, Bremner JB, Rattanaphani V. An adsorption study of Al(iii) ions onto chitosan. J Hazard Mater. 2007;148(1–2):185–91.10.1016/j.jhazmat.2007.02.024Search in Google Scholar PubMed

[34] Haider S, Binagag FF, Haider A, Al-Masry WA. Electrospun oxime-grafted-polyacrylonitrile nanofiber membrane and its application to the adsorption of dyes. J Polym Res. 2014;21(371):1–13.10.1007/s10965-014-0371-1Search in Google Scholar

[35] Jain R, Sharma P, Sikarwar S. Kinetics and isotherm analysis of tropaeoline 000 adsorption onto unsaturated polyester resin (UPR): a non-carbon adsorbent. Environ Sci Pollut R. 2013;20(3):1493–502.10.1007/s11356-012-0994-xSearch in Google Scholar PubMed

[36] Wang WQ, Zhang YT, Yue YR, Zhang B, Liu JD. Adsorption of Cu(ii) from aqueous solution by porous Mn3[Co(CN)6]2·nH2O nanospheres. Sep Sci Technol. 2014;49(2):232–40.10.1080/01496395.2013.835824Search in Google Scholar

[37] Crini G, Badot PM. Application of chitosan, a natural aminopolysaccharide, for dye removal from aqueous solutions by adsorption processes using batch studies: a review of recent literature. Prog Polym Sci. 2008;33(4):399–447.10.1016/j.progpolymsci.2007.11.001Search in Google Scholar

Received: 2020-02-14
Revised: 2020-03-30
Accepted: 2020-04-19
Published Online: 2020-06-02

© 2020 Hongya Li et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Obituary for Prof. Dr. Jun-ichi Yoshida
  2. Regular Articles
  3. Optimization of microwave-assisted manganese leaching from electrolyte manganese residue
  4. Crustacean shell bio-refining to chitin by natural deep eutectic solvents
  5. The kinetics of the extraction of caffeine from guarana seed under the action of ultrasonic field with simultaneous cooling
  6. Biocomposite scaffold preparation from hydroxyapatite extracted from waste bovine bone
  7. A simple room temperature-static bioreactor for effective synthesis of hexyl acetate
  8. Biofabrication of zinc oxide nanoparticles, characterization and cytotoxicity against pediatric leukemia cell lines
  9. Efficient synthesis of palladium nanoparticles using guar gum as stabilizer and their applications as catalyst in reduction reactions and degradation of azo dyes
  10. Isolation of biosurfactant producing bacteria from Potwar oil fields: Effect of non-fossil fuel based carbon sources
  11. Green synthesis, characterization and photocatalytic applications of silver nanoparticles using Diospyros lotus
  12. Dielectric properties and microwave heating behavior of neutral leaching residues from zinc metallurgy in the microwave field
  13. Green synthesis and stabilization of silver nanoparticles using Lysimachia foenum-graecum Hance extract and their antibacterial activity
  14. Microwave-induced heating behavior of Y-TZP ceramics under multiphysics system
  15. Synthesis and catalytic properties of nickel salts of Keggin-type heteropolyacids embedded metal-organic framework hybrid nanocatalyst
  16. Preparation and properties of hydrogel based on sawdust cellulose for environmentally friendly slow release fertilizers
  17. Structural characterization, antioxidant and cytotoxic effects of iron nanoparticles synthesized using Asphodelus aestivus Brot. aqueous extract
  18. Phase transformation involved in the reduction process of magnesium oxide in calcined dolomite by ferrosilicon with additive of aluminum
  19. Green synthesis of TiO2 nanoparticles from Syzygium cumini extract for photo-catalytic removal of lead (Pb) in explosive industrial wastewater
  20. The study on the influence of oxidation degree and temperature on the viscosity of biodiesel
  21. Prepare a catalyst consist of rare earth minerals to denitrate via NH3-SCR
  22. Bacterial nanobiotic potential
  23. Green synthesis and characterization of carboxymethyl guar gum: Application in textile printing technology
  24. Potential of adsorbents from agricultural wastes as alternative fillers in mixed matrix membrane for gas separation: A review
  25. Bactericidal and cytotoxic properties of green synthesized nanosilver using Rosmarinus officinalis leaves
  26. Synthesis of biomass-supported CuNi zero-valent nanoparticles through wetness co-impregnation method for the removal of carcinogenic dyes and nitroarene
  27. Synthesis of 2,2′-dibenzoylaminodiphenyl disulfide based on Aspen Plus simulation and the development of green synthesis processes
  28. Catalytic performance of the biosynthesized AgNps from Bistorta amplexicaule: antifungal, bactericidal, and reduction of carcinogenic 4-nitrophenol
  29. Optical and antimicrobial properties of silver nanoparticles synthesized via green route using honey
  30. Adsorption of l-α-glycerophosphocholine on ion-exchange resin: Equilibrium, kinetic, and thermodynamic studies
  31. Microwave-assisted green synthesis of silver nanoparticles using dried extracts of Chlorella vulgaris and antibacterial activity studies
  32. Preparation of graphene oxide/chitosan complex and its adsorption properties for heavy metal ions
  33. Green synthesis of metal and metal oxide nanoparticles from plant leaf extracts and their applications: A review
  34. Synthesis, characterization, and electrochemical properties of carbon nanotubes used as cathode materials for Al–air batteries from a renewable source of water hyacinth
  35. Optimization of medium–low-grade phosphorus rock carbothermal reduction process by response surface methodology
  36. The study of rod-shaped TiO2 composite material in the protection of stone cultural relics
  37. Eco-friendly synthesis of AuNPs for cutaneous wound-healing applications in nursing care after surgery
  38. Green approach in fabrication of photocatalytic, antimicrobial, and antioxidant zinc oxide nanoparticles – hydrothermal synthesis using clove hydroalcoholic extract and optimization of the process
  39. Green synthesis: Proposed mechanism and factors influencing the synthesis of platinum nanoparticles
  40. Green synthesis of 3-(1-naphthyl), 4-methyl-3-(1-naphthyl) coumarins and 3-phenylcoumarins using dual-frequency ultrasonication
  41. Optimization for removal efficiency of fluoride using La(iii)–Al(iii)-activated carbon modified by chemical route
  42. In vitro biological activity of Hydroclathrus clathratus and its use as an extracellular bioreductant for silver nanoparticle formation
  43. Evaluation of saponin-rich/poor leaf extract-mediated silver nanoparticles and their antifungal capacity
  44. Propylene carbonate synthesis from propylene oxide and CO2 over Ga-Silicate-1 catalyst
  45. Environmentally benevolent synthesis and characterization of silver nanoparticles using Olea ferruginea Royle for antibacterial and antioxidant activities
  46. Eco-synthesis and characterization of titanium nanoparticles: Testing its cytotoxicity and antibacterial effects
  47. A novel biofabrication of gold nanoparticles using Erythrina senegalensis leaf extract and their ameliorative effect on mycoplasmal pneumonia for treating lung infection in nursing care
  48. Phytosynthesis of selenium nanoparticles using the costus extract for bactericidal application against foodborne pathogens
  49. Temperature effects on electrospun chitosan nanofibers
  50. An electrochemical method to investigate the effects of compound composition on gold dissolution in thiosulfate solution
  51. Trillium govanianum Wall. Ex. Royle rhizomes extract-medicated silver nanoparticles and their antimicrobial activity
  52. In vitro bactericidal, antidiabetic, cytotoxic, anticoagulant, and hemolytic effect of green-synthesized silver nanoparticles using Allium sativum clove extract incubated at various temperatures
  53. The green synthesis of N-hydroxyethyl-substituted 1,2,3,4-tetrahydroquinolines with acidic ionic liquid as catalyst
  54. Effect of KMnO4 on catalytic combustion performance of semi-coke
  55. Removal of Congo red and malachite green from aqueous solution using heterogeneous Ag/ZnCo-ZIF catalyst in the presence of hydrogen peroxide
  56. Nucleotide-based green synthesis of lanthanide coordination polymers for tunable white-light emission
  57. Determination of life cycle GHG emission factor for paper products of Vietnam
  58. Parabolic trough solar collectors: A general overview of technology, industrial applications, energy market, modeling, and standards
  59. Structural characteristics of plant cell wall elucidated by solution-state 2D NMR spectroscopy with an optimized procedure
  60. Sustainable utilization of a converter slagging agent prepared by converter precipitator dust and oxide scale
  61. Efficacy of chitosan silver nanoparticles from shrimp-shell wastes against major mosquito vectors of public health importance
  62. Effectiveness of six different methods in green synthesis of selenium nanoparticles using propolis extract: Screening and characterization
  63. Characterizations and analysis of the antioxidant, antimicrobial, and dye reduction ability of green synthesized silver nanoparticles
  64. Foliar applications of bio-fabricated selenium nanoparticles to improve the growth of wheat plants under drought stress
  65. Green synthesis of silver nanoparticles from Valeriana jatamansi shoots extract and its antimicrobial activity
  66. Characterization and biological activities of synthesized zinc oxide nanoparticles using the extract of Acantholimon serotinum
  67. Effect of calcination temperature on rare earth tailing catalysts for catalytic methane combustion
  68. Enhanced diuretic action of furosemide by complexation with β-cyclodextrin in the presence of sodium lauryl sulfate
  69. Development of chitosan/agar-silver nanoparticles-coated paper for antibacterial application
  70. Preparation, characterization, and catalytic performance of Pd–Ni/AC bimetallic nano-catalysts
  71. Acid red G dye removal from aqueous solutions by porous ceramsite produced from solid wastes: Batch and fixed-bed studies
  72. Review Articles
  73. Recent advances in the catalytic applications of GO/rGO for green organic synthesis
Downloaded on 8.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2020-0032/html
Scroll to top button