

Synthesis and Absorption Properties of Leaf-like Nd2Cu2O4+δ via a Coordination Complex Method with [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O Precursor
English
Synthesis and Absorption Properties of Leaf-like Nd2Cu2O4+δ via a Coordination Complex Method with [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O Precursor
-
1. INTRODUCTION
Until the end of the nineteenth century, all the pigments were extracted from natural substances, like flowers and roots. However, with the development of modern synthetic chemistry and chemical industry, natural dyes have been replaced by synthetic ones because they can be manufactured on a large scale, which is one the most important factors in water pollution. The artificial dyes are so stable that it is difficult to degrade in nature, which may significantly impact the health of human beings[1, 2]. MG is an organic dye widely used in cotton, silk, wool, leather and paper industries, and also a drug used to treat parasites, fungal and bacterial infections[3, 4]. Despite its extensive usage, MG dye has caused several health hazards as carcinogenesis, mutagenesis, teratogenesis and respiratory toxicity agent[5, 6]. Therefore, the removal of dyes from effluents is very important. Up to now, several treatment methods including adsorption[2-4, 7], photodegradation[8] and biological treatment[9] have been developed to remove dyes from effluents. Among them, adsorption technology grows rapidly because of its relatively low cost and effect in dye removal.
Up to now, various adsorbents have been reported to eliminate MG from effluents, such as porous carbon[10], zeolites[11], mesoporous materials[12, 13], some metal-organic frameworks (MOFs)[14] and so on. However, the results of these studies showed limited adsorption efficiencies which restrict their applications in large scale effluents treatment. Recently, coordination complexes consisting of metal ions or clusters and organic ligands have been used as alternative precursors for the synthesis of nanostructure adsorbents[15–18], and some of these materials exhibit good adsorption properties of organic dyes from aqueous solution. However, most of them are focused on transition metal complexes, and the design and control over lanthanide-transition bimetal complexes are relatively less.
In this study, Nd2Cu2O4+δ samples were prepared by a coordination-complex method (CCM) taking [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O (1) as a precursor to yield particles with large special surface areas when compared to those prepared by simple solution method (SSM). The as-prepared Nd2Cu2O4+δ particles were characterized by using XRD, SEM, HRTEM, XPS, and BET methods. The adsorption properties are studied in detail, showing Nd2Cu2O4+δ samples exhibit selective adsorption towards MG with significant Qm (maximum adsorption capacity) values reaching up to 1.55 g/g at room temperature.
2. MATERIALS AND METHODS
2.1 Synthesis of Nd2Cu2O4+δ
The [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O (1) was prepared according to the method published in the literature[19, 20] as the CCM precursor. A mixture of Cu(OAc)2·4H2O (0.5 mmol, 127 mg), Nd(NO3)3·6H2O (0.5 mmol, 219 mg), 3,4-pdc (1 mmol, 167 mg) and triethylamine (0.2 mL) was dissolved in a mixture of water-methanol at the volume ratio of 1:1. The solution was stirred for 3 hours then filtered off and allowed to stand undisturbed. After about 5 days, blue block crystals suitable for X-ray analysis were obtained by filtration, washing with ethyl ether and drying in air. The obtained single-crystal precursor was then calcined at different temperature for 1 hour under N2 atmosphere, followed by 0.5 h in air to yield Nd2Cu2O4+δ.
The SSM samples were synthesized according to the method reported in the literature[19, 20]. Under constant stirring, Cu(OAc)2·4H2O (7.9 mmol, 2 g) and Nd(NO3)3·6H2O (7.9 mmol, 3.46 g) were dissolved in 50 mL distilled water at the stoichiometric proportions. After 1 h, the solution was heated to a gel and then calcined at 800 ℃ for 1 h to yield Nd2Cu2O4+δ.
2.2 Characterization
Single-crystal X-ray diffraction data of the coordination precursor [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O (1) were collected at 100 K by a Rigaku Saturn CCD diffractometer, equipped with graphite-monochromated Mo-Kα radiation with a radiation wavelength of 0.71073 Å using an ω-φ scan mode. The structure was solved by direct methods and refined with a full-matrix least-squares technique based on F2 using the SHELXS-97 and SHELXL-97 programs[21]. Anisotropic thermal parameters were assigned to all non-hydrogen atoms. Crystallographic data and structural refinement parameters are listed in Table S1 (the Supporting Information). The selected bond lengths and bond angles for 1 are given in Table S2.
XRD patterns were examined on a Rigaku D/Max-RB X-ray diffractometer with CuKα irradiation with a scan speed (2θ) of 0.05 °/s from 10 to 90°. Powder morphologies were performed by SEM (Zeiss Supra 55) and HRTEM (FEI Tecnai F30). The specific surface areas of the as-prepared samples were characterized by N2 adsorption/desorption experiments at 77 K with a Builder SSA-4300 instrument. The XPS measurements were measured on a PHI 5000 C ESCA System with Mg K source operating at 14.0 kV and 25 mA.
2.3 Adsorption experiments
Adsorption properties were carried out to evaluate the adsorption capacity of Nd2Cu2O4+δ towards MG. Batch adsorption experiments were carried out in conical flasks containing 1000 mL of 0.4 g/L MG aqueous/ethanol mixture solution at 1:1 volume ratio and various adsorbent doses (0.03~0.07 g) at different temperature (298, 313, and 338 K) for 7 days. Once the equilibrium was established, 5 mL sample solution was extracted by a micropipette and the adsorbent particles were removed by centrifugation. The concentration of residual MG in the sample was determined by UV-Vis spectrophotometer (T9) at maximum absorbance of 618 nm. The efficiencies of the adsorbent in sequestration of MG dye, expressed in terms of adsorption capacity at the equilibrium concentration, qe (mg/g), were calculated according to Eq. (1):
$q_e=\frac{\left(C_0-C_e\right) \times V}{m}$ (1) where C0 (mg/L) and Ce (mg/L) are the initial and equilibrium concentrations of MG in the solution, respectively, V (L) is the initial solution volume, and m (g) is the used dry adsorbent mass.
3. RESULTS AND DISCUSSION
3.1 Characterization
The X-ray diffraction analysis reveals that [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O (1) crystallizes in the monoclinic system with space group P21/n. The asymmetric unit in 1 (Fig. 1) consists of one crystallographically independent Nd3+, one Cu2+, two 3,4-pdc ligand anions, one acetate ion and five coordinated water molecules. Nd3+ is coordinated by two carboxylic oxygen atoms from two 3,4-pdc ligands, three oxygen atoms from two acetate ions and four water molecules, and the coordination geometry for the nine-coordinated Nd3+ can be described as a distorted tricapped trigonal prism. Two N atoms and two carboxylic oxygen atoms located at diagonal positions respectively and one water molecule lies at the axial position, finishing the square pyramidal coordination sphere of Cu2+. The bond lengths of Nd–O (3,4-pdc) are 2.372(4) and 2.471(4) Å, and the bond lengths of Nd–O (acetate) are 2.637(4), 2.471(4) and 2.542(5) Å. The bond lengths of Cu–O (carboxylic) are 1.982(4) and 1.962(3) Å, and those of Cu–N are 2.016(4) and 2.005(4) Å. The Cu–O (water) is 2.304(4) Å, which is longer than other Cu–O bonds, due to the Jahn-Teller effect. Compared to the similar complexes reported in Ref. [19], the axial Cu–O bond length of complex 1 is longer, indicating that the coordination ability of the water molecule is weaker. Both of 3,4-pdc ligands adopt a tri-dentate mode to coordinate with one Nd3+ ion and two Cu2+ ions (Fig. 1b). Cu2+ ions are bridged by 3,4-pdc ligands, giving rise to a (4,4)-connected 2D lattice plane with a regular rhombus of about 9 × 9 Å2 size (defined by the distances between Cu2+ ions), as shown in Fig. 1c. Two Nd3+ ions are bridged by acetate ligands, adopting a μ2-η2: η1 bridging mode, forming a dinuclear Nd(Ⅲ) unit (Fig. 1d). The 2D lattice planes are further connected by the dinuclear Nd(Ⅲ) units, giving rise to a 3D sandwich-like framework (Fig. 1e). It should be noted that the structure has a large porosity along the a and b directions. A PLATON program analysis suggests that there is approximately 38% of the crystal volume accessible to the solvents after taking off the lattice water molecules.
Figure 1
Figure 1. (a) Molecular structure of 1, (b) Coordination mode of 3,4-pdc ligand, (c) 2D lattice plane constructed by Cu2+ and 3,4-pdc, (d) Connection mode of the 2D lattice plane and the dinuclear Nd(Ⅲ) unit, (e) 3D sandwich-like framework viewed along the a directionXRD was used to investigate phase structures of the as-prepared Nd2Cu2O4+δ particles, and Fig. 2 shows the XRD patterns of the sample prepared at different temperature, indicating that Nd2Cu2O4+δ was formed above 800 ℃. All the observed diffraction peaks can be indexed to the standard data of Nd2CuO4 with PDF card no 80-1644[22-25]. When below 800 ℃, the CuO and Nd2O3 crystallites were initially formed due to their low activation energies and relatively high energies of formation[26]. No obvious impurity phase was detected above 800 ℃. The XRD of SSM shows that the samples of SSM contain more Nd2O3 impurities under the same conditions (Fig. S1). Compared to SSM, the sample of CCM was purer, which could be conjectured by the fact that the metal ions were evenly distributed in the coordination precursor. However, the stoichiometric ratio of the coordination precursor is 1:1, which is different with the stoichiometric ratio of Nd2CuO4, indicating that Cu2+ ions were doped into the host lattice of Nd3+.
Figure 2
SEM and TEM are performed to observe the morphology of Nd2Cu2O4+δ particles. Fig. 3 shows the results of SEM and TEM investigations for samples prepared at 900 ℃. Figs. 3a and 3b show the results of SEM investigations for Nd2Cu2O4+δ prepared by CCM, which can be described as a leaf-like nanosheet with a width of dozens of microns and a thickness of 50~80 nm. Most leaves are stacked together, but the layer structure is still clearly observed. It is suggested that the coordination precursor may be lamellar division firstly during the decomposition process and the organic matter volatilized at high temperature[26]. From the high magnification picture, we can see droplet particles on the lamellae, which are characterized by SEM elemental analysis (Figs. 3c and 3d), indicating that the distribution of elements in the droplet is also Nd: Cu = 1:1 phase[23]. These results are consistent with the Nd/Cu stoichiometric ratio and show that the proportion of elements in the configuration of Nd2Cu2O4+δ is indeed 1:1 and distributes evenly.
Figure 3
The clear lattice spacing detected by HRTEM
(Fig. 3e) suggested that Nd2Cu2O4+δ particles were highly crystalline in nature with single crystalline structure[27]. The interplanar crystal spacing is 0.374 nm, indicating that the side of Nd2Cu2O4+δ term is (101) crystal face[28]. Fig. 3f shows the morphology of SEM investigations for Nd2Cu2O4+δ prepared by SSM, which can be described as an inhomogeneous massive particle of 0.5~4 um. Compared with the samples prepared by CCM, the SSM one has a smaller surface area and is not suitable for the adsorbent.
To further characterize the porosity of Nd2Cu2O4+δ prepared by CCM, nitrogen adsorption-desorption experiments were performed at 77 K. The nitrogen adsorption-desorption isotherms and pore size distribution are shown in Fig. 4. As evidenced in Fig. 4, the Nd2Cu2O4+δ of CCM showed type Ⅲ isotherm profiles according to IUPAC classification, indicating weak adsorbent-adsorbate interaction[29, 30]. Although the leaf-like nanosheets prepared by CCM do not have channels deep into the matrix, the distance between slices is relatively close. Therefore, the pore distribution statistics show that there are aperture distributions in the range of 10~100 nm, which are consistent with the SEM observations. The calculated BET surface area is 17.9 m2/g.
Figure 4
The surface chemical composition and elemental states of Nd2CuO4 adsorbents are investigated by XPS. Fig. 5a presents the XPS survey spectrum of Nd2CuO4 of CCM, showing that the sample contains Nd, Cu and O. The high-resolution XPS spectra of Nd, Cu and O were attentively deconvoluted by considering spin-orbital coupling. The high-resolution XPS spectra of Nd 3d are shown in Fig. 5b. The peaks of 3d3/2 and 3d5/2 observed respectively at 1004.8 and 982.0 eV suggest the presence of Nd ion with a formal charge of +3[30]. Meanwhile, one little shoulder peak, observed at 1000.1 eV, might be attributed to the satellite peak of Nd ion with +3~δ1 charge, which possibly implies that a part of Nd ions occupies the O4-coordinated sites (Fig. 1). As shown in Fig. 5c, the Cu 2p XPS of CCM showed core level of Cu 2p spectral region with two spin-orbital doublets. The main peaks presented Cu 2p1/2 at 953.8 eV and Cu 2p3/2 at 934 eV with an energy difference of about 20 eV, which could be attributed to Cu ion in CuO4 group with a formal charge of +2[31]. The second doublet with binding energies at 962.9 and 943.2 eV could be ascribed to the emission from Cu 2p1/2 and Cu 2p3/2 core levels of Cu atoms with more positive charges, denoted as +2 + δ2, suggesting that an inequivalent Cu ion locates at an 8-coordinated position[31, 32]. Fig. 5d shows two different valences of O at 528.9 and 531.3 eV (more positive)[33, 34], respectively, indicating that there are two kinds of non-equivalent O atoms: one (528.9 eV) is related to 4-coordinated O species[35] and the other (531.3 eV) to 6-coordinated O species. The XPS profiles of Nd2CuO4 by SSM are almost the same (Fig. S2), which implies that the Cu and Nd ions cloud occupy the 4- and 6-coordinated sites in the A2BO4 structure, respectively, and have a mixed valence state. This further demonstrates that the A2BO4-type compounds with perovskite structure have good solubility to metal ions[31].
Figure 5
3.2 Maximum adsorption capacity of Nd2Cu2O4+δ for MG
The adsorption capabilities toward MG of the as-prepared Nd2Cu2O4+δ were evaluated at 298, 318 and 338 K. In order to investigate the adsorption capacity of Nd2Cu2O4+δ, a series of absorbent doses of CCM samples (0.03, 0.04, 0.05, 0.06 and 0.07 g) were employed to study the equilibrium isotherm, and the concentration of residual MG in the sample was determined by UV-Vis spectrophotometer (T9) at maximum absorbance, 618 nm, as shown in Fig. 6.
Figure 6
The adsorption isotherms play an important role in determining the adsorption ability of MG over Nd2Cu2O4+δ. The adsorption isotherms were investigated by Langmuir and Freundlich isotherm models, which could be expressed as follows:
$\frac{1}{q_e}=\frac{1}{K^\theta Q_m} \times \frac{1}{C_e}+\frac{1}{Q_m}$ (2) $\ln q_e=\ln K_F+\frac{1}{n} \ln C_e$ (3) The correlation between measuring equilibrium concentrations of MG, Ce, and their corresponding equilibrium adsorption capacity (qe) can induce a linear plot of 1/Ce versus 1/qe for Langmuir model (Eq. 2) or a linear plot of lnCe versus lnqe for Freundlich model, Eq. (3)[31]. The best fitting yields the maximum adsorption amount (Qm) and Kθ (Langmuir constants) or KF (Freundlich constants) and n (intensity factor). Furthermore, the values of related parameters and R2 of CCM samples are provided in Table 1. Fig. 6d suggests that the adsorption experimental data of dye over Nd2Cu2O4+δ are well described using Langmuir isotherm model with R2 being closer to 1. The Qm of Nd2Cu2O4+δ estimated by Langmuir model reached 1.55 g/g at 298 K with the equilibrium constant (Kθ) equal to 6395 L/mol. To the best of our knowledge, the maximum adsorption amount of Nd2Cu2O4+δ towards MG is relatively larger, which is comparable to the reported value in the literature[35, 36]. As temperature increased, the maximum adsorption value of Nd2Cu2O4+δ decreased to 1.25 g/g at 338 K, accompanying the equilibrium constant falling to 2254 L/mol. The thermodynamic parameters can be obtained from the fitting of lnKθ versus T−1 according to Eq. (4), as shown in Fig. 6f.
$\ln K^\theta=\frac{-\Delta_r G_m{ }^\theta}{R T}=-\frac{\Delta_r H_m^\theta}{R} \times \frac{1}{T}+\frac{\Delta_r S_m{ }^\theta}{R}$ (4) Table 1
T/K Langmuir Freundlich Qm (g/g) KL (L/mol) R2 KF (L/mol) 1/n R2 298 1.55 6395 0.9968 2099 0.339 0.8494 318 1.35 4323 0.9951 1282 0.363 0.7756 338 1.25 2254 0.9994 1357 0.563 0.7860 The standard Gibbs free energy change (ΔrGmθ), standard enthalpy (ΔrHmθ) change and standard entropy change (ΔrSmθ) for adsorption of 1 mol MG are calculated to be –21.80, –21.68 and 0.51 J/mol·K, respectively. The negative value of ΔrGmθ indicates that the adsorption reaction is spontaneous. The negative value of ΔrHmθ further confirms the decrease of equilibrium constant with increasing the temperature. The positive value of ΔrSmθ may imply that the surface of the adsorbent is covered initially by water molecules and the adsorbed MG molecule occupies a larger area on the surface.
4. CONCLUSION
Nd2Cu2O4+δ adsorbents were successfully synthesized via CCMs taking [NdCu(3,4-pdc)2(OAc)(H2O)5]·6.5H2O as the coordination precursor. Compared to the particles prepared by SSM, the sample of CCM was purer and showed leaf-like morphology composed of nanosheets with an average thickness of 50~80 nm and a BET surface area up to 17.9 m2/g. XRD, SEM mapping, and XPS measurements showed that Nd2Cu2O4+δ is A2BO4 -type compound with Cu2+ ions occupying the A site of the host lattice. Adsorption capabilities were investigated, and Nd2Cu2O4+δ samples exhibit selective adsorption towards malachite green (MG) with significant Qm (maximum adsorption capacity) values reaching up 1.55 g/g at room temperature. The fitting of the isotherms at different temperature estimated the following thermodynamic parameters: ΔrGmθ –21.8 kJ/mol, ΔrHmθ –21.68 kJ/mol, and ΔrSmθ 0.51 J/mol·K, indicating the adsorption reaction is a spontaneous process. Therefore, the Nd2Cu2O4+δ may be utilized as an ideal candidate for applications in environmental and separation science.
-
-
[1]
He, Y. H.; Chen, Z. X.; Xu, J. J.; Wu, Y.; Xiao, G. C. Solvothermal synthesis of ZnIn2S4 by alcohol solvents and visible light photocatalytic activity on selective oxidation and dye degradation. Chin. J. Struct. Chem. 2018, 37, 753−762.
-
[2]
Zhang, Y. L.; Yang, J.; Yu, X. J. Preparation, characterization, and adsorption-photocatalytic activity of nano TiO2 embedded in diatomite synthesis materials. Rare Met. 2017, 36, 987−991. doi: 10.1007/s12598-014-0290-7
-
[3]
Xue, H.; Ding, N.; Lai, S. W.; Chen, Q. H.; Liu, X. P.; Qian, Q. R. Rapid microwave-assisted hydrothermal synthesis of Bi0.76Sb1.24S3 and its application in the photocatalytic degradation of pollutants by visible light irradiation. Chin. J. Struct. Chem. 2017, 36, 59−65.
-
[4]
Azad, F. N.; Ghaedi, M.; Dashtian, K.; Hajati, S.; Goudarzi, A.; Jamshidi, M. Enhanced simultaneous removal of malachite green and safranin O by ZnO nanorod-loaded activated carbon: modeling, optimization and adsorption isotherms. New J. Chem. 2015, 39, 7998−8005. doi: 10.1039/C5NJ01281C
-
[5]
Asfaram, A.; Ghaedi, M.; Goudarzi, A.; Soylak, M.; Langroodi, M. S. Magnetic nanoparticle based dispersive micro-solid-phase extraction for the determination of malachite green in water samples: optimized experimental design. New J. Chem. 2015, 39, 9813−9823. doi: 10.1039/C5NJ01730K
-
[6]
Huang, P.; Zhao, P.; Dai, X.; Hou, X.; Zhao, L.; Liang, N. Trace determination of antibacterial pharmaceuticals in fishes by microwave-assisted extraction and solidphase purification combined with dispersive liquid-liquid microextraction followed by ultra-high performance liquid chromatography-tandem mass spectrometry. J. Chromatogr. B. 2016, 1011, 136−144. doi: 10.1016/j.jchromb.2015.12.059
-
[7]
Gao, Q.; Luo, J.; Wang, X. Y.; Gao, C. X.; Ge, X. Q. Novel hollow alpha-Fe2O3 nanofibers via electrospinning for dye adsorption. Nano. Res. Lett. 2015, 10, 176−8. doi: 10.1186/s11671-015-0874-7
-
[8]
Wu, Y. T.; Li, M. L.; Yuan, J.; Wang, X. F. Synthesis and characterizations of metastable Bi2SiO5 powders with a synergistic effect of adsorption and photocatalysis. Appl. Phys. A 2017, 123, 543−10. doi: 10.1007/s00339-017-1144-6
-
[9]
Huang, D.; Hu, C.; Zeng, G.; Cheng, M.; Xu, P.; Gong, X.; Wang, R.; Xue, W. Combination of Fenton processes and biotreatment for wastewater treatment and soil remediation. Sci. Total Environ. 2017, 574, 1599−1610. doi: 10.1016/j.scitotenv.2016.08.199
-
[10]
Yu, M.; Li, J.; Wang, L. KOH-activated carbon aerogels derived from sodium carboxymethyl cellulose for high-performance supercapacitors and dye adsorption. Chem. Eng. J. 2017, 310, 300−306. doi: 10.1016/j.cej.2016.10.121
-
[11]
Shaw, R.; Sharma, R.; Tiwari, S.; Tiwari, S. K. Surface engineered zeolite: an active interface for rapid adsorption and degradation of toxic contaminants in water. ACS Appl. Mater. Inter. 2016, 8, 12520−12527. doi: 10.1021/acsami.6b01754
-
[12]
Tian, Y.; Liu, P.; Wang, X. F.; Lin, H. S. Adsorption of malachite green from aqueous solutions onto ordered mesoporous carbons. Chem. Eng. J. 2011, 171, 1263−1269. doi: 10.1016/j.cej.2011.05.040
-
[13]
Xu, R.; Jia, M.; Zhang, Y. L.; Li, F. T. Sorption of malachite green on vinyl-modified mesoporous poly(acrylic acid)/SiO2 composite nanofiber membranes. Micro. Meso. Mater. 2012, 149, 111−118. doi: 10.1016/j.micromeso.2011.08.024
-
[14]
Hasan, Z.; Jhung, S. H. Removal of hazardous organics from water using metal-organic frameworks (MOFs): plausible mechanisms for selective adsorptions. J. Hazard. Mater. 2015, 283, 329−339. doi: 10.1016/j.jhazmat.2014.09.046
-
[15]
Lee, H. J.; Cho, W.; Lim, E.; Oh, M. One-pot synthesis of magnetic particle-embedded porous carbon composites from metal-organic frameworks and their sorption properties. Chem. Commun. 2014, 50, 5476−5479. doi: 10.1039/c4cc01914h
-
[16]
Zhang, C.; Ye, F.; Shen, S.; Xiong, Y.; Su, L.; Zhao, S. From metal-organic frameworks to magnetic nanostructured porous carbon composites: towards highly efficient dye removal and degradation. RSC Adv. 2015, 5, 8228−8235. doi: 10.1039/C4RA15942J
-
[17]
Wang, L.; Ke, F.; Zhu, J. Metal-organic gel templated synthesis of magnetic porous carbon for highly efficient removal of organic dyes. Dalton Trans. 2016, 45, 4541−4547. doi: 10.1039/C5DT04260G
-
[18]
Yang, Y.; Zhang, Y.; Sun, C. J.; Li, X.; Zhang, W.; Ma, X.; Ren, Y.; Zhang, X. Heterobimetallic metal-organic framework as a precursor to prepare a nickel/nanoporous carbon composite catalyst for 4-nitrophenol reduction. ChemCatChem. 2014, 6, 3084−3090. doi: 10.1002/cctc.201402607
-
[19]
Liu, X. W.; Guo, R.; Liu, H.; Yu, Y. Q.; Qi, X. W.; Xu, J. Y.; Xie, C. Z. Two series of novel 3D potentially porous heterometallic Cu−Ln coordination frameworks assembled by 3,4-pyridinedicarboxylic acid with different topologies and channels: syntheses, structures, luminescence and magnetic properties. RSC Adv. 2015, 5, 15059−15068. doi: 10.1039/C4RA13533D
-
[20]
You, J. H.; Guo, Y. Z.; Guo, R.; Liu, X. W. A review of visible light-active photocatalysts for water disinfection: Features and prospects. Chem. Eng. J. 2019, 372, 624−641. doi: 10.1016/j.cej.2019.04.192
-
[21]
Sheldrick, G. M. SHELXS-97 and SHELXL-97, Program for Crystal Structure Solution and Refinement. Göttingen University, Germany 1997.
-
[22]
Yang, H. Q.; Liu, S. W.; Cao, L. H.; Jiang, S. H.; Huo, H. Q. Superlithiation of non-conductive polyimide toward high-performance lithium-ion batteries. J. Mater. Chem. A 2018, 6, 21216−21224. doi: 10.1039/C8TA05109G
-
[23]
Jiang, S. H.; Uch, B.; Agarwal, S.; Greiner, A. Ultralight, thermally insulating, compressible polyimide fiber assembled sponges. ACS Appl. Mater. Inter. 2017, 9, 32308−32315. doi: 10.1021/acsami.7b11045
-
[24]
Duan, G. G.; Liu, S. W.; Jiang, S. H.; Hou, H. Q. High-performance polyamide-imide films and electrospun aligned nanofibers from an amidecontaining diamine. J. Mater. Sci. 2019, 54, 6719−6727. doi: 10.1007/s10853-019-03326-w
-
[25]
Xu, H. B.; Jiang, S. H.; Ding, C. H.; Zhu, Y. M.; Li, J. J.; Hou, H. Q. High strength and high breaking load of single electrospun polyimide microfiber from water soluble precursor. Mater. Lett. 2017, 201, 82−84. doi: 10.1016/j.matlet.2017.05.019
-
[26]
Liu, X. W.; Wang, R. C.; Ni, Z. Y.; Zhou, W. L.; Du, Y. C.; Ye, Z. Q.; Guo, R. Facile synthesis and selective adsorption properties of Sm2CuO4 for malachite green: kinetics, thermodynamics and DFT studies. J. Alloy. Compd. 2018, 743, 17−25. doi: 10.1016/j.jallcom.2018.01.320
-
[27]
Lv, D.; Wang, R. X.; Tang, G. S.; Mou, Z. P.; Lei, J. D.; Han, J. Q.; Smedt, S. D.; Xiong, R. H.; Huang, C. B. Ecofriendly electrospun membranes loaded with visible-light responding nanoparticles for multifunctional usages: highly efficient air filtration, dye scavenging, and bactericidal activity. ACS Appl. Mater. Inter. 2019, 11, 12880−12889. doi: 10.1021/acsami.9b01508
-
[28]
Ding, Q. Q.; Xu, X. W.; Yue, Y. Y.; Mei, C. T.; Huang, C. B.; Jiang, S. H.; Wu, Q. L.; Han, J. Q. Nanocellulose-mediated electroconductive self-healing hydrogels with high strength, plasticity, viscoelasticity, stretchability, and biocompatibility toward multifunctional applications. ACS Appl. Mater. Inter. 2018, 10, 27987−28002. doi: 10.1021/acsami.8b09656
-
[29]
Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.; Rouquerol, J.; Siemieniewska, T. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity. Pure Appl. Chem. 1985, 57, 603−619. doi: 10.1351/pac198557040603
-
[30]
Yousefi, T.; Torab-Mostaedi, M.; Aghaei, A.; Ghasemi-Mobtaker, H. Facile synthesis, morphology and structure of Dy2O3 nanoparticles through electrochemical precipitation. Rare Met. 2016, 35, 637−642. doi: 10.1007/s12598-015-0448-y
-
[31]
Biesinger, M. C.; Lau, L. W. M.; Gerson, A. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Sc, Ti, V, Cu and Zn. Appl. Surf. Sci. 2010, 257, 887−898. doi: 10.1016/j.apsusc.2010.07.086
-
[32]
Jiang, J. Z.; Wang, Y. L.; Wang, C. Y.; Bai, L. M.; Li, X.; Li, Y. B. Magnesium hydroxide whisker modified via in situ copolymerization of n-butyl acrylate and maleic anhydride. Rare Met. 2017, 36, 997−1002. doi: 10.1007/s12598-016-0826-0
-
[33]
Guo, R.; Yan, A. G.; Xu, J. J.; Xu, B. T.; Li, T. T.; Liu, X. W.; Yi, T. F.; Luo, S. H. Effects of morphology on the visible-light-driven photocatalytic and bactericidal properties of BiVO4/CdS heterojunctions: A discussion on photocatalysis mechanism. J. Alloy. Compd. 2020, 817, 153245−12.
-
[34]
Liu, X. W.; Ni, Z. Y.; He, Y.; Su, N.; Guo, R.; Wang, Q.; Yi, T. F. Ultrasound-assisted two-step water-bath synthesis of g-C3N4/BiOBr composites: visible light-driven photocatalysis, sterilization, and reaction mechanism. New. J. Chem. 2019, 43, 8711−8721. doi: 10.1039/C9NJ01398A
-
[35]
Tian, Y.; Liu, P.; Wang, X. F.; Lin, H. S. Adsorption of malachite green from aqueous solutions onto ordered mesoporous carbons. Chem. Eng. J. 2011, 171, 1263−1269. doi: 10.1016/j.cej.2011.05.040
-
[36]
Deshpande, P. A.; Polisetti, S.; Madras, G. Rapid synthesis of ultrahigh adsorption capacity zirconia by a solution combustion technique. Langmuir. 2011, 27, 3578−3587. doi: 10.1021/la104674k
-
[1]
-
Table 1. Partial Fitting Results Obtained from the Maximum Adsorption Capacity Experiment of CCM
T/K Langmuir Freundlich Qm (g/g) KL (L/mol) R2 KF (L/mol) 1/n R2 298 1.55 6395 0.9968 2099 0.339 0.8494 318 1.35 4323 0.9951 1282 0.363 0.7756 338 1.25 2254 0.9994 1357 0.563 0.7860 -

计量
- PDF下载量: 1
- 文章访问数: 525
- HTML全文浏览量: 9