Jinsong Heab,
Jiangdong Daiab,
Atian Xieab,
Sujun Tianab,
Zhongshuai Changab,
Yongsheng Yan*ab and
Pengwei Huo*ab
aSchool of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China. E-mail: yys@mail.ujs.edu.cn; huopw@mail.ujs.edu.cn; Fax: +86-0511-88791800; Tel: +86-0511-88790683
bInstitute of Green Chemistry and Chemical Technology, Jiangsu University, Zhenjiang 212013, China
First published on 17th August 2016
The development of practicable and retrievable adsorbents with high adsorption capacity is a technical imperative for water treatment. Herein, we reported a new convenient macroscopic granular adsorbent for the removal of tetracycline (TC) from water by immobilizing porous carbons (PCs) which were obtained via one-step in situ pyrolysis from ethylenediaminetetraacetic acid dipotassium salt dihydrate (EDTA-2K·2H2O) into carboxymethylcellulose sodium (CMCS) gel beads utilising molecular cross-linking. A remarkable similarity can be observed between multivariant gel beads and EDTA-2K·2H2O derived porous carbons (EPCs) according to the characterization results. The adsorption performance was evaluated using batch adsorption studies of TC in aqueous solution: the kinetic results could be fit well by a pseudo-second-order model and intraparticle diffusion was treated as the rate-controlling step; equilibrium adsorption data fitted well to the Langmuir adsorption isotherm yielding a maximum adsorption capacity of 136.9 mg g−1 at 298 K. Importantly, these results indicate that the as-prepared gel beads could be used as facile adsorbents in pharmaceutical wastewater treatment.
PCs can be prepared using template synthesis and activation processes. Template synthesis has been employed frequently for the synthesis of uniform and/or interconnected porous carbon materials, and despite the the shortcomings of this approach, such as preparation complexity and high cost, it has been successfully applied in energy3 and other particular fields.4 In nature, there is no rigorous demand for homogeneous porosity when PCs is used as an absorbent; activation process are generally adopted for the production of adsorbents. However, the risk and complexity incurred in physical/chemical activation hinder the practical production of PCs. In order to overcome this predicament, direct carbonization, which does not involve an activation process, was employed. The selection of precursor is crucial to obtain PCs with prominent morphology and chemical structure. Several kinds of precursor, such as phenolic resin,5 tetraphenylborate,6 polyvinylidene fluoride,7 coordination polymers,8 zeolitic imidazolate frameworks,9 and metal–organic frameworks1 have been tested so far; nevertheless, there is still an increasing demand for the development of appropriate and attractive precursors.
EDTA-2K·2H2O is a nitrogen-containing organic potassium salt with the formula of C10H14K2N2O8·2H2O, which consists of carboxyl groups, potassium, alkyl chains and amidocyanogen. Taking this as the foundation, in situ doping of carboxyl groups and potassium ions into precursors can realize the construction of homogeneous and abundant pore structures due to the release of a mixed gas of H2, CO, CO2 and potassium vapor as a superb pore-forming agent. In addition, substantial production can be guaranteed by the plenitudinous alkyl chains. Moreover, functional groups derived from oxygen and nitrogen containing groups are required for high efficiency PCs in environmental applications. In summary, EDTA-2K·2H2O is able to act as a remarkable precursor for the preparation of PCs.
Even so, it is noted that the separation of these porous carbon after adsorption requires lengthy coagulation and/or high speed centrifugation; and these will cause secondary pollution due to the particle loss caused by the potential release of adsorbent into the environment, which limits their applications in large scale wastewater treatment processes.10 To overcome this problem, many researchers focused their attentions on gel beads, as they could be entrapped. On one hand, complex separation and recollection are no longer necessary when using gel beads as adsorbent. On the other hand, gels obtained from alginate11 and chitosan12 are widely used for wastewater treatment due to their large scale availability, acceptable cost, biodegradability and biocompatibility. Apart from these two raw materials, we surprisingly find that carboxymethylcellulose sodium (CMCS) can also be used as cross-linking agent; furthermore, in the presence of trivalent cations, such as Fe3+, it forms hydrogels via intermolecular crosslinking.
CMCS is a derivative of natural cellulose. Cellulose, the most important skeletal component in plants, is the most abundant natural polymer and has been considered an almost inexhaustible source of raw material for the increasing demand for biocompatible products.13 In this way, it has an extensively expanded application potential when compared with alginate and chitosan. Moreover, CMCS also has considerable fascinating properties, such as low cost, biodegradability, non-toxicity, and high stability. As a promising candidate, it can also be widely used as a biosorbents due to the presence of carboxylate groups along the polymeric chains.14
In this paper, we present a very simple method to prepare PCs with a high specific surface area, huge volume and abundant chemical groups using EDTA-2K·2H2O as precursor via a one-step in situ pyrolysis process. Furthermore, we use CMCS as cross-linking agent to aggregate PCs for the solution separation and successful recollection of adsorbent from the adsorbed matrix solution. The physical and chemical properties of as-prepared compound gel beads were characterized by scanning electron microscopy (SEM), surface area analysis, X-ray photoelectron spectroscopy (XPS), Fourier transform-infrared (FTIR) spectroscopy, X-ray diffraction (XRD) and Raman spectroscopy. Herein, tetracycline (TC) was chosen as adsorbate to investigate the removal efficiency of antibiotics from water, as it is the most used antibiotic as animal feed additive and dominant source of antibiotic resistant genes.15 The potential of sorption was evaluated by adsorption kinetics, equilibrium and thermodynamic studies. The developed valuable and facile route for preparation of macroscopic spherical EDTA-2K·2H2O derived porous carbon–carboxymethylcellulose sodium (EPCs@CMCS) gel beads may provide a positive route for promising environmental remediation.
Fig. 1 Digital photo of EPCs@CMCS gel beads appearing with different sizes in the wet and dry state. |
The adsorption equilibrium experiments were carried out with various initial concentrations varying from 25 to 200 mg L−1 of the solution at pH 6.0, and were kept in a constant temperature water bath for 24 h to reach an apparent sorption equilibrium based on the results of adsorption kinetics.
In order to prove the prepared materials have potential in real sample applications, the effects of co-existing anion species and humic acid (HA) on the adsorption of TC onto gel beads was investigated. The experiments were conducted in TC solution (100 mg L−1) containing Cl−, NO3−, SO24−, PO34− (0.01 M), and HA (10 mg L−1) at 298 K for 24 h. In the pH effect experiments, the initial pH values of the TC solutions (100 mg L−1) were adjusted in the range of 3.0–8.0 using dilute NaOH or HCl solutions.
All adsorption experiments were repeated at least three times to ensure accuracy of the obtained data. The amount of adsorption at time t (Qt, mg g−1) was calculated by eqn (1).
(1) |
SEM images of EPCs@CMCS gel beads are demonstrated in Fig. 2. The surface morphology is uneven and ridgy (Fig. 2a), which can be elucidated by the asynchronous shrinking of EPCs in the gel bead mixture, which occurred during freeze drying. The appearance of conspicuous cracks on the surface of the EPCs@CMCS gel beads was perhaps due to excess implantation of EPCs in CMCS gel beads leading to an overloading of the crosslinker burden. The crosslinked inner nature is shown in Fig. 2b; the crosslinker skeleton structure is composed of many caves and surfaces covered with EPCs. When compared with SEM images of CMCS gel beads in Fig. S1 (see the ESI†), the visible changes in the internal skeleton structure and the surface morphology of the gel beads could be discriminated easily. The interconnected network structure and internal surface prolongation of EPCs@CMCS gel beads facilitated the attachment of adsorbate to the adsorption active sites.
It is worth noting that these well-porous carbon materials can be easily produced in only one-step in situ pyrolysis by using EDTA-2K·2H2O as precursor. The excellent porosity of the EPCs is elucidated in Table 1; they exhibit a high BET specific surface area (SBET) (i.e. 2916 m2 g−1) and a large pore volume (VP) (i.e. 1.63 cm3 g−1), moreover, micropores are dominant in the EPCs. In the process of self-activation, carboxyl groups and potassium carboxylate converted into K2CO3 and CO2 below 400 °C firstly, then carbon was consumed by the reaction of carbon and H2O with the emission of H2 and CO (eqn (2)); sequentially, fractional CO reacted with H2O with the emission of CO2 and H2 (eqn (3)).
C + H2O → CO + H2 | (2) |
CO + H2O → CO2 + H2 | (3) |
Samples | SBET/m2 g−1 | VP/cm3 g−1 | Smicro/m2 g−1 | Vmicro/cm3 g−1 | daver/nm |
---|---|---|---|---|---|
a SBET: BET specific surface area; VP: total pore volume; Smicro: micropore surface area; Vmicro: micropore volume; daver: average pore diameter. | |||||
EPCs@CMCS gel beads | 1501 | 0.77 | 739.2 | 0.31 | 2.05 |
EPCs | 2916 | 1.63 | 2156 | 1.01 | 1.68 |
When temperatures rose to 700 °C, significant decomposition occurred on the as-formed K2CO3 into CO2 and K2O. Furthermore, the K compounds also can be reduced by carbon to produce metallic K vapor at temperatures over 700 °C (eqn (5) and (6)).18
K2CO3 → K2O + CO2 | (4) |
K2CO3 + 2C → 2K + 3CO | (5) |
C + K2O → 2K + CO | (6) |
The homogeneous and abundant pore structure could be realized due to the release of a mixed gas of H2, CO, CO2 and metallic K as an active pore-forming agent.19
Many researchers have proved that micropores are major positions for organic contaminant adsorption onto carbon composite materials, therefore, an abundant porosity of adsorption materials could contribute to highly efficient organic matter adsorption.16,20 The N2 adsorption–desorption isotherms of EPCs@CMCS gel beads and EPCs exhibit type I isotherms (according to the IUPAC classification) with a steep increase region at low relative pressure (P/Po < 0.1) and a plateau region with a parallel relative pressure axis in the range of P/Po > 0.1 (Fig. 3a), which correspond to micropore filling and multilayer adsorption on the surface, respectively.21 The pore distribution of EPCs@CMCS gel beads in Fig. 3b clearly indicates that micropores are more directly dominant; the mesopore distribution of the products is in the range of 0.5–3.0 nm. Table 1 gives the comparison results of porous textural characterization. From the results, although SBET and VP decreased after saturation of CMCS in EPCs, marvellously, the pore size distribution and microporous proportion remained.
Fig. 3 N2 adsorption–desorption isotherms of EPCs@CMCS gel beads and EPCs (a), pore size distribution (b), high-resolution XPS spectra of C 1s (c) and O 1s (d) peaks of EPCs@CMCS gel beads. |
The nature of most external surface functionalities of EPCs@CMCS gel beads was also studied by XPS, the spectra of C 1s and O 1s photoelectrons are shown in Fig. 3c and d. The high resolution C 1s photoelectron spectrum for the raw hydrochar sample included four common signals, which can be attributed to the aliphatic/aromatic carbon groups (C1, C–H/C, 284.5 eV), hydroxyl groups (C2, C–OH, 286.0 eV), carbonyl groups (C3, CO, 287.1 eV), and carboxylic/ester/lactone groups (C4, C–OOH, 288.7 eV).22,23 The O 1s spectra can be deconvoluted into three groups: ketone, lactone, carbonyl, and quinone groups (O1, CO, 531.1 eV); carbonyl oxygen atoms in esters, amides, carboxylic anhydrides and oxygen atoms in hydroxyls or ethers (O2, C–OH and/or C–O–C, 532.4 eV); and carboxyl hydroxyl or perhaps some adsorbed water (O3, O–CO/H2O, 533.4 eV).24 The C 1s and O 1s XPS spectra of EPCs are shown in Fig. S2.† It was observed that the surface functionality species were identical with EPCs@CMCS gel beads, however, the content of oxygen-containing functional groups was lower.
FT-IR spectra are presented in Fig. 4a; for CMCS gel beads, the stretching vibration peak of O–H appeared at 3441 cm−1, the peak at 2908 cm−1 is assigned to the aromatic and aliphatic C–H groups. The band at 1733 cm−1 may be attributed to CO stretching, peaks at 1601 and 1418 cm−1 can be contributed to carboxyl group OC–O asymmetric stretching vibrations and symmetrical stretching vibrations, respectively. The peaks at 1111, 1067 and 1032 cm−1 related to C–O stretching in ether, carboxyl acids and alcohols.25 Pure carbon material EPCs have finite characteristic absorption bands at 1032–1111 cm−1 (C–O stretching) and 878 cm−1 (C–O–H vibration).26 After impregnation of EPCs in CMCS beads, the peak at 1601 cm−1 disappeared and a new peak appeared at 1560 cm−1 which was assigned to the formation of cross-linking. In addition, the peak of O–H attenuated due to the reduction of the number of hydroxyls per unit. The absorption bands of EPCs@CMCS gel beads maintain characteristic peaks of CMCS gel beads and are more similar to EPCs.
Fig. 4 FT-IR spectra (a), Raman spectra (b) and XRD patterns (c) of EPCs, CMCS gel beads and EPCs@CMCS gel beads. |
In the Raman spectra (Fig. 4b), two prominent peaks at 1356 and 1582 cm−1 correspond to the D band and the G band, the D band derives from the stretching vibration of the defect sp3 carbon atoms, whereas the G band originates from the stretching vibration of sp2 carbon networks. The integrated intensity ratio ID/IG was widely used to characterise the degree of defects in graphitic composites;27 there was minimal increase of ID/IG ratio after immobilization of CMCS into EPCs, indicating the introduction of defects.
The XRD patterns (Fig. 4c) revealed the amorphous state. The 100 diffraction peaks located at about 43.2° suggested an ordered hexagonal carbon structure; peaks around 24.4° correspond to 002 diffraction peaks indicating the degree of stacking order of the layered carbon structure.23 As shown in Fig. 4c, after immobilization of EPCs into CMCS gel beads, the 2θ value of 002 planes increased from 23.6° to 24.8°, on the basis of Bragg’s law, nλ = 2dsinθ, where λ is the wavelength (1.54 Å for Cu Kα), indicating the reduction of interlayer spacing (d002) from 0.38 nm to 0.36 nm, which may be attributed to the interaction of the interstitial atoms.28 The weakness of the characteristic 100 diffraction peak in EPCs@CMCS gel beads reflected that the graphitization degree reduced after immobilization.
Fig. 5 Effect of solution pH (a) and co-existing anions and HA (b) on the removal of TC by EPCs@CMCS gel beads. |
Pseudo-first-order: ln(Q1 − Qt) = lnQ1 − k1t | (7) |
(8) |
Intraparticle diffusion: Qt = kidt1/2 + C | (9) |
The parameters of the determination coefficient (R2) and normalized standard deviation (ΔQ, %) eqn (10) were evaluated to define which model best describes the experimental data.
(10) |
The adsorption process was originally rapid and then slow, after which it remained stagnant, due to the fact that the available vacant surface sites had decreased with the prolongation of adsorption. The process of EPCs@CMCS gel beads TC removal lasted for a long time as equilibrium was obtained around 8 h when the initial concentration was 100 mg L−1; the higher the initial TC concentration, the longer the adsorption time required (Fig. 6a). Compared with powder adsorbent applications for TC removal, such as activated carbon22 or activated sludge,33 EPCs@CMCS gel beads took a longer time to reach equilibrium. This occurs as a result of the lower effective contact surface and slow-rate transmission channel of massive beads.
The fitting parameters of pseudo-first-order and pseudo-second-order kinetics are listed in Table 2 and the linear plots are presented in Fig. S3.† The pseudo-second-order model showed a higher regression coefficient and a lower normalized standard deviation compared with the pseudo-first-order model. This result indicates that the better representation of adsorption kinetics is the pseudo-second-order model, and its Qe,cal values agree well with Qe,exp values. The pseudo-second-order kinetic model is based on the experimental information of solid-phase sorption, which means that the combined action of physical and chemical absorption is occurring.
Co/mg L−1 | Pseudo-first-order | Pseudo-second-order | ||||||
---|---|---|---|---|---|---|---|---|
k1/min−1 | Q1,cal/mg g−1 | ΔQe (%) | R2 | k2/10−3 g mg−1 min−1 | Q2,cal/mg g−1 | ΔQe (%) | R2 | |
50 | 0.0952 | 29.41 | 21.62 | 0.7816 | 11.4 | 49.02 | 12.73 | 0.9969 |
75 | 0.0921 | 48.63 | 21.56 | 0.8124 | 5.82 | 74.07 | 13.46 | 0.9943 |
100 | 0.0798 | 64.02 | 27.23 | 0.8248 | 5.41 | 93.46 | 16.51 | 0.9937 |
Generally, the adsorption process onto an adsorbent can be divided into three stages. The first stage is external diffusion, in which the adsorbates move from the bulk solution to the external surface of the adsorbent. The second stage is intraparticle diffusion, where the adsorbates diffuse further to the adsorption sites within the adsorbent. The last stage is adsorbing, in this stage the adsorbates are adsorbed at the active sites on the adsorbent, which is a fast step and is usually negligible.34 The transportation of adsorbate from solution phase to the surface of the adsorbent particles may be controlled either by one or more steps; an intraparticle diffusion model can be used to elucidate the diffusion mechanism well. According to Fig. 6b, the plots of Qt to t1/2 exhibited a piecewise-linear pattern with two slopes. The first inclined stage is attributed to a combination of rapid adsorption of TC on the impacted surface via cracks distributed on the external surface of EPCs@CMCS gel beads and intraparticle diffusion. The last portion is the final adsorbing equilibrium stage where intraparticle diffusion starts to be retarded due to the decrease of the free TC concentration. It is essential for Qt versus t1/2 plots to go through the origin if the intraparticle diffusion is the sole rate-limiting step; as shown in Fig. 6b and Table S1,† the intraparticle diffusion can be regarded approximately as a rate-controlling step with different initial TC concentrations, which is a relatively good explanation of the time-lapse adsorption rate.
(11) |
(12) |
(13) |
Adsorption isotherm non-linear fitting results are described in Fig. 7 and fitting parameters are listed in Table 3. The correlation coefficient of fitting the linear form to the Langmuir model was much higher than that of two other models under all temperature conditions; this meant that the Langmuir model was more suitable to imitate the adsorption process, which indicated monolayer coverage of TC on the available surface of gel beads (the maximum capacity based on Langmuir was 136.9 mg g−1 at 298 K). Meanwhile, the other two models fit reasonably, showing correlations of isotherms in the order: Langmuir > Temkin > Freundlich. The fitting result of the Temkin model indicated that there was electrostatic interaction during the adsorption process. The essential features of the Langmuir isotherms RL indicate the possibility of the adsorption process being irreversible (RL = 0), favorable (0 < RL < 1), linear (RL = 1), or unfavorable (RL > 1).38
(14) |
T/K | Langmuir | Freundlich | Temkin | |||||||
---|---|---|---|---|---|---|---|---|---|---|
Qm/mg g−1 | KL/L mg−1 | RL | R2 | KF/(mg g−1) (L mg−1)1/n | n | R2 | bT | KT/L g−1 | R2 | |
288 | 104.2 | 0.08 | 0.33 | 0.9763 | 14.69 | 2.25 | 0.7681 | 105.46 | 1.35 | 0.8667 |
298 | 136.9 | 0.13 | 0.23 | 0.9995 | 24.95 | 2.39 | 0.9202 | 89.47 | 1.57 | 0.9806 |
308 | 149.2 | 0.16 | 0.20 | 0.9991 | 28.89 | 2.36 | 0.9331 | 83.55 | 1.96 | 0.9876 |
As the value of RL fell between 0 and 1, thisindicated that TC adsorption on EPCs@CMCS gel beads was favorable. In addition, the parameter n derived from the Freundlich model also indicated the favorability of sorption.2,39
As compared with other previously reported adsorbents listed in Table 4, the cost-effective EPCs@CMCS gel beads were not only were provided with an appreciable adsorption performance for TC from an aquatic environment, but also were able to be recollected conveniently. In addition, the secondary pollution of particle loss caused by the potential release of adsorbent into the environment could be effectively avoided. It is suggested that the prepared gel beads have great potential as an easily retrievable adsorbent for antibiotics removal.
Adsorbent | Qm/mg g−1 | Adsorption conditions | Ref. |
---|---|---|---|
Graphene oxide | 370 | pH = 3.6, T = 298 K | 26 |
Carbon nanotubes | 269.5 | T = 293 K | 39 |
Amino-Fe functionalized mesoporous silica | 96.9 | pH = 5.0 ± 0.1, T = 298 K, 0.01 M NaCl | 44 |
Alkali bio-char | 58.8 | Ambient pH, T = 303 K | 45 |
Illite | 32 | pH = 5–6 | 46 |
Magnetic porous carbon | 29 | T = 303 K | 40 |
CMCS gel beads | 9.47 | pH = 6.0, T = 298 K | This work |
EPCs@CMCS gel beads | 136.9 | pH = 6.0, T = 298 K | This work |
ΔG° = −RTlnKo | (15) |
Ko represents the ability of the adsorbent to retain the adsorbate and the extent of movement of the adsorbate within the solution; values of Ko are obtained by plotting a straight line of ln(Cs/Ce) as a function of Cs. Cs is the amount of TC adsorbed per unit gram of the adsorbent (mmol g−1) and Ce is the equilibrium concentration of the TC (mmol mL−1).
The enthalpy change (ΔH°) and entropy change (ΔS°) of adsorption were estimated from eqn (16).
(16) |
The adsorption capacities obviously increased with the rise of temperature, indicating that the adsorption is an endothermic reaction. Its spontaneous and thermodynamically favorable nature was also explained by the negative values of ΔG° and positive value of ΔH° (Table 5). The positive ΔS° indicated the increase of randomness at the solid/liquid interface during the adsorption process.40
T/K | lnKo | ΔG°/kJ mol−1 | ΔH°/kJ mol−1 | ΔS°/J mol−1 K−1 |
---|---|---|---|---|
288 | 2.194 | −5.619 | 6.251 | 40.09 |
298 | 2.334 | −5.976 | ||
308 | 2.363 | −6.050 |
The negative values of ΔG° and positive value of ΔH° confirm that the adsorption processes was endothermic, matching well with the adsorption being more favorable at higher temperature. Generally, ΔG° for physisorption is between −20 and 0 kJ mol−1 and for chemisorption is between −80 and −400 kJ mol−1. Furthermore, Kara et al.41 suggested that value of ΔH° for physical adsorption was smaller than 40 kJ mol−1. Herein, both the values of ΔG° and ΔH° indicated that the adsorption processes was mainly a physical adsorption. Moreover, the positive values of ΔH° evidenced the increment of randomness at the solid/solution interface during the adsorption process.42
Furthermore, chemisorption is also involved in the process; although the attenuation of peak intensity after immobilization occurred, the remaining 100 diffraction peaks in the XRD patterns (Fig. 4c) indicated the graphitization of EPCs@CMCS gel beads. TC in possession of conjugated enone can act as a commendable π-electron acceptor, due to the strong electron-withdrawing ability of the ketone group, and therefore interact strongly with the short stacks of graphite-like sheets (π-electron-donors) arranged in a disordered form of the EPCs@CMCS gel beads via π–π EDA interactions.16 In addition, electrostatic interactions and hydrophobic interaction may be also involved in the process on the basis of the solution pH results and Temkin model fitting results.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra14877h |
This journal is © The Royal Society of Chemistry 2016 |