1 s2.0 S0143720822000948Cd Main

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Dyes and Pigments 202 (2022) 110172

Contents lists available at ScienceDirect

Dyes and Pigments


journal homepage: www.elsevier.com/locate/dyepig

Three Cd(II) coordination polymers containing phenylenediacetate isomers:


Luminescence sensing and adsorption antibiotics performance in water
Mei-Li Zhang a, **, Xue-Ying Lu a, Ye Bai a, Yi-Xia Ren a, Ji-Jiang Wang a, Xiao-Gang Yang b, *
a
Department of Chemistry and Chemical Engineering, Laboratory of New Energy & New Function Materials, Yan′ an University, Yan′ an, China
b
College of Chemistry and Chemical Engineering, Luoyang Normal University, Henan Province Function-Oriented Porous Materials Key Laboratory, Luoyang, China

A R T I C L E I N F O A B S T R A C T

Keywords: Three new Cd(II) coordination polymers (CPs) [Cd(opda) (tib)]⋅H2O (1), [Cd4(mpda)4(tib)2(dib)2]⋅7H2O (2), [Cd
Coordination polymers (CPs) (ppda) (tib)]⋅4H2O (3), have been synthesized by using isomeric phenylenediacetates (H2pda) and 1,3,5-tris(1-
Antibiotics imidazolyl)benzene (tib) ligand under hydrothermal condition. Single crystal X-ray structures showed that CP 1
Luminescent sensors
constitutes 3,5-connected 2-nodal net with a three-dimensional (3D) structure, while CP 2 forms an ring 1D
Adsorption performance
chain, CP 3 has a interpenetrated 2D structure. The structural diversity of these CPs is mainly attributed to the
different coordination modes and isomer conformations of the flexible phenylenediacetic acids. Three CPs are
highly selective and sensitive luminescent sensors toward tetracycline (TEC). Compared with CPs 2 and 3, CP 1
has higher quenching percentage and adsorption capacities (64.493 mg g− 1) for TEC due to its large specific
surface area (171.72 m2 g− 1) and pore volume (0.31 cm3 g− 1). The main adsorption mechanisms are phys­
isorption and chemisorption, including pore filling effect and π-π interaction. This study provides new insights
into the design of 3D CPs for removing pollutants from industrial wastewaters, with the potential to mitigate
health risks to humans and natural ecosystems.

1. Introduction adsorption of contaminants removal in aqueous. Many CPs and their


modified materials have been used to remove antibiotics from waste­
The increasing use of antibiotics in many fields brings out negative water, such as Fe3O4-ES/ZIF-8 [24], MOF-5 [25], [Zn6(ID­
environmental effects [1–5]. However, removal of antibiotics from C)4(OH)2(Hprz)2] [26] and MIL-101(Cr)@GO [27]. These CPs not only
wastewater is difficult due to its changeful occurrence state, low exhibit high adsorption but can also be used in sustainable recycling.
biodegradability, and complexity of molecular structures [6]. Therefore, Since the CPs have showed great potential in adsorptive removal of
the high-efficiency removal of antibiotics from aquatic system has antibiotics from aquatic environment, it is urgent to synthesis CPs-based
already become a important topic. Currently, many methods for the adsorbent with other properties. As far as we know, the aqueous
removal of antibiotics have been developed and enhanced, including adsorption of tetracycline (TEC) onto the Cd-CPs has not been
oxidation [7], reverse osmosis [8], ion exchange [9], and adsorption investigated.
[10]. Comparatively speaking, adsorption techniques are an efficient Within this context, we herein expect Cd-CPs with porous property to
and economical approach for removing antibiotics, involving lower be an advanced adsorbent for highly effective removal of antibiotics
costs, easier implementation, and reduced pollution [11,12]. from aqueous solutions (Scheme 1). In this work, we are focusing on the
Adsorbents are generally porous materials with large specific surface influence of the coordination chemistry of phenylenediacetate isomers
areas, such as activated carbon [13], carbon nanotubes [14], graphene with different coordination groups, conformations, and flexibility. Three
oxide [15], biochar [16], porous silica [17], molecularly imprinted new 3D/1D/2D Cd(II) coordination polymers (CPs 1–3) have been
polymers [18], porous resins, and metal oxides [19]. CPs have some synthesized and structurally characterized. Furthermore, their lumi­
unique advantages, such as ultrahigh BET surface area, high/tunable nescent properties and sensing behaviors have also been investigated in
porosity, various pore functionality and structures, open metal sites, and water solution. The as-synthesized samples of these CPs could be used
so on [20–23]. Therefor, CPs with porous property have been applied in directly as TEC sensor because of their sensitive and specific

* Corresponding author.
** Corresponding author.
E-mail address: [email protected] (M.-L. Zhang).

https://doi.org/10.1016/j.dyepig.2022.110172
Received 12 January 2022; Received in revised form 28 January 2022; Accepted 7 February 2022
Available online 15 February 2022
0143-7208/© 2022 Elsevier Ltd. All rights reserved.
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

fluorescence response. More interestingly, CP 1 has the best fluorescence 2.2. Syntheses of title CPs
quenching effect on TEC, which may be because CP 1 has a 3D porous
network. Specifically, the quenching mechanism of CPs 1–3 to detect [Cd(opda) (tib)]⋅H2O (1) A mixture of H2opda (0.0210 g, 0.1
TEC have been investigated in detail. UV–vis experiments showed that mmol), Cd(NO3)2⋅4H2O (0.0308 g, 0.1 mmol), symtiyb (0.0276 g, 0.1
the absorption spectrum of TEC had the biggest overlap with the emis­ mmol) and NaOH (8.0 mg, 0.2 mmol) were added to water (12 mL) in a
sion spectrum of the as-obtained fluorescent matter (CPs 1–3), which is 25 mL Teflon-lined stainless steel vessel. The mixture was heated at
favorable to the absorption of emission light. The theoretical calculation 160 ◦ C for 72 h. After the reactive mixture was slowly cooled to room
confirmed that the LUMO energy level of TEC is lower than that of CPs temperature, colourless block crystal of 1 was obtained. Elemental
1–3, which is available to the donor− acceptor electron transfer. More­ analysis calcd. (%) for C25H22N6O5Cd: C 50.15, H 3.68, N 14.03; found
over, we investigated the adsorption performance of three CPs for (%): C 50.03, H 3.43, N 13.89. IR (KBr pellet, cm− 1): 3442 m, 3131 w,
removal of TEC. CP 1 exhibited the best adsorption capacities for TEC 2278 w, 1622 s, 1559 s, 1519 m, 1367 s, 1073 m, 1012 w, 753 m, 654 w.
due to its large specific surface area and pore volume. The main [Cd4(mpda)4(tib)2(dib)2]⋅7H2O (2) CP 2 was synthesized by a
adsorption mechanisms were physisorption and chemisorption, procedure similar to that of CP 1, except H2mpda replaced H2opda.
including pore filling effect and π-π interaction. Furthermore, the Herein, dib was generated by the decomposition of tib. Colourless block
experiment influencing factors, i.e., reaction time, pH, ionic strength crystal of CP 2 was obtained. Elemental analysis calcd. (%) for
and reusability were analyzed synthetically. These superior properties of C94H92Cd4N22O23: C 48.05, H 3.92, N 13.12; found (%): C 47.89, H 3.81,
CP 1 indicated that it could be applied to treatment of antibiotic N 12.97. IR (KBr pellet, cm− 1): 3455 m, 3141 w, 2379 w, 1608 s, 1577 s,
wastewater. Obviously, this study provides a new choice for the detec­ 1519 m, 1396 m, 1352 s, 1112 m, 950 m, 853 m, 654 s.
tion and adsorption of TEC antibiotics in aqueous solution, and has a Cd(ppda) (tib)]⋅4H2O (3) CP 3 was synthesized by a procedure
broad application prospect in water quality monitoring. similar to that of CP 1, except H2ppda replaced H2opda. Colourless block
crystal of CP 3 was obtained. Elemental analysis calcd. (%) for
2. Experimental C25H28N6O8Cd: C 45.95, H 4.29, N 12.87; found (%): C 45.91, H 4.12, N
12.81. IR (KBr pellet, cm− 1): 3448 m, 3137 w, 2344 w, 1618 s, 1567 s,
2.1. Materials and general methods 1512 s, 1396 m, 1360 s, 1073 m, 1012 w, 744 m, 649 w.

All the starting reagents and solvents were commercially available 2.3. Preparation of fluorescent probe detection experiment
and used as received without further purification. The hydrothermal
reaction was performed in a 25 mL Teflon-lined stainless steel autoclave CPs 1–3 (3 mg, respectively) were placed in 10 mL water, treated by
under autogenous pressure. Elemental analyses for C, H, and N were ultrasonication for 1 h. The aqueous solution of CPs 1–3 were allowed to
carried out on a Flash 2000 organic elemental analyzer. Thermal stand for 1 days to form a stable suspension. Eight common antibiotic
gravimetric analyses (TGA) were carried out on a SDT Q600 thermog­ (roxithromycin (ROX), tetracycline (TEC), chloramphenicol (CHL),
ravimetric analyzer with a heating rate of 10 ◦ C/min under a N2 at­ gentanicin sulfate (GEN), penicilin sodium (PEN), lincomycin hydro­
mosphere. Powder X-ray diffraction (PXRD) measurements were chloride (LIN), cefixime (CEF), azithromycin (AZI)) stock solutions were
performed on a Bruker D8-ADVANCE X-ray diffractometer with Cu Kα prepared using ethanol/DMF and water at a concentration of 1 mM,
radiation (λ = 1.5418 Å). Solid state and liquid fluorescence spectrum respectively. The prepared aqueous solutions of CPs 1–3 were used for
was measured by fluorescence spectrophotometer (F-7100) at room the initial luminescence spectra and fluorescence experiments, adding
temperature. UV–vis absorption were carried out on a UV–vis antibiotic to the prepared aqueous solutions of CPs 1–3 each time. The
spectrophotometer(UV-2700). Density functional theory is used to CP 1 was washed with ultrapure water and ethanol, and the colorless
calculate the energy theory. The fluorescence decay curves and quantum and transparent crystals obtained were dried under vacuum at 80 ◦ C for
yield were tested on the FLSP920 transient steady-state fluorescence 8 h to prepare for the adsorption of antibiotics.
spectrometer.

Scheme 1. Three-dimensional porous CPs was used for fluorescence response and adsorption removal of tetracycline (TEC).

2
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

2.4. Preparation of adsorption antibiotic experiment Table 1


Crystallographic data for CPs 1–3.
The whole adsorption experiment adopts the method of batch Complex 1 2 3
adsorption. By dissolving 200 mg of tetracycline in a 1L ultrapure water
Formula C25H22N6O5Cd C94H92Cd4N22O23 C25H28N6O8Cd
to prepare 200 mg/L of the tetracycline stock solution. The solutions of Mass 598.88 2347.49 652.93
different concentrations required for the experiment can be obtained by Temperature/K 296 296 296
continuously diluting the stock solution with ultrapure water. All Crystal system Monoclinic Monoclinic Monoclinic
adsorption experiments were performed by adding 5 mg of adsorbent to Space group Cc P21/c P21/c
a (Å) 13.332(12) 13.3259(19) 11.8035(6)
20 mL of tetracycline solution in a 50 mL erlenmeyer flask, and per­ b (Å) 10.142(12) 12.4005(17) 12.2750(7)
formed on a THZ-82A constant temperature water bath shaker with a c (Å) 18.291(18) 29.842(4) 19.7528(11)
rotation speed of 200 rpm. Adjust the solution to the target pH by α (◦ ) 90 90 90
dropping a negligible volume of sulfuric acid or sodium hydroxide, and β (◦ ) 91.39(2) 100.859(2) 103.5590(10)
γ (◦ ) 90 90 90
react for 24 h in the dark. After the antibiotics are adsorbed in the water,
V (Å3) 2472(4) 4842.8(12) 2782.18(30)
the solution is solid-liquid separated at a speed of 8000 revolutions per Z 4 2 4
minute using a centrifuge, and filtered through a disposable poly­ Dc (g cm− 3) 1.609 1.610 1.559
vinylidene fluoride filter with a pore diameter of 0.45 μM. Determine the μ (mm− 1) 0.932 0.951 0.843
concentration of the remaining tetracycline in the solution by measuring Rint 0.0275 0.0531 0.0250
Goof 1.133 1.038 1.106
the absorbance of tetracycline at 357 nm using an ultraviolet spectro­ R1a/wR2b [I >2σ(I)] 0.0275/0.0706 0.0531/0.1232 0.0367/0.1011
photometer. The standard curve is determined by using the standard R1a/wR2b (all data) 0.0311/0.0843 0.1102/0.1606 0.0437/0.1139
solution of tetracycline in the concentration range 5–60 mg/L (Fig. S1). a
R1 = Σ||Fo| – |Fc||/Σ|Fo|, wR2 = [Σw(Fo2 – Fc2)2/Σw(Fo2)2]1/2.
The adsorption capacity of the adsorbent can be calculated by the
following eq (1):
3. Results and discussion
qe=(C0–Ce)V/m (1)
3.1. Crystal structure description
Where C0(mg L− 1) and Ce(mg L− 1) represent the initial concentration
and equilibrium concentration of the tetracycline solution, m(g) is the
3.1.1. Structure of [Cd(opda) (tib)]⋅H2O (1)
mass of the adsorbent, and V(L) is the volume of the tetracycline
Single-crystal X-ray diffraction analysis reveals that CP 1 crystallizes
solution.
in the monoclinic crystal system, Cc space group. The Cd1 atom lies in
Conduct adsorption kinetic experiments to study the adsorption rate
distorted octahedral coordination geometry, and coordinated with three
and determine the type of adsorption, add 5 mg of adsorbent to 20 mL of
nitrogen atoms of three different tib ligands and three oxygen atoms
a tetracycline solution with an initial concentration of 40 mg/L for
from two different opda2− ligands (Fig. 1a). The tib ligand acting as a
adsorption experiments, and measure the tetracycline concentration of
three connected node joins the Cd(II) ions to form a 2D extended
the solution at specific time intervals (5min-24h). The adsorption
structure, in which the metal ions are also three connected (Fig. 1b). The
experiment was carried out at three different temperatures (25, 35,
two carboxylic groups of trans-conformation opda2− anion, displaying
45 ◦ C) at pH 7.0 to confirm the maximum adsorption capacity and
single-dentate and double-dentate chelation coordination modes
thermodynamic properties of the material. For adsorption isotherm
(Fig. S2a). These adjacent layers are joined together by trans-confor­
studies, 5 mg of adsorbent was added to 20 mL of tetracycline solution
mation opda ligands to form a 3D porous network (Fig. 1c). Overall, the
with a concentration range of 5–200 mg/L, and the adsorption experi­
metal atoms act as five connected nodes. The simplified 3D net structure
ment was carried out at 25 ◦ C and pH 7.0. By adding negligible volume
of CP 1 can be described in the topological representation as 3,5-con­
of sulfuric acid and sodium hydroxide to the tetracycline solution to
nected 2-nodal net.
adjust the pH value of the solution, the effect of pH on the adsorption of
tetracycline on the material was studied in the range of pH 3–9. Finally,
3.1.2. Structure of [Cd4(mpda)4(tib)2(dib)2]⋅7H2O (2)
ethanol is used as the eluent to wash the material loaded with TEC drug,
Single crystal X-ray structure analysis reveals that the asymmetric
and the washed material is used for the reusability experiment of
unit of CP 2 contains four Cd(II) atoms, four fully deprotonated mpda2−
adsorption.
ligand, two tib and two dib ligands. As shown in Fig. 2a, each Cd(II)
atom in CP 2 is six-coordinated or seven-coordinated. In CP 2, two
2.5. X-ray crystal structure
carboxy groups of mpda in CP 2 adopt the μ1–η1:η1 chelate coordination
mode and pint to the same side of the benzene ring plan. Thus mpda2−
X-ray single-crystal diffraction data for CPs 1–3 were collected at
ligand has a cis-conformation and can be seen as a “V” type building
room temperature, using Bruker Smart 1000 CCD area-detector
block (Fig. S2b). Four unique Cd atoms are bridged by four mpda ligands
diffractometer with Mo-Kα radiation (λ = 0.71073 Å) by ω scan mode.
to form a [Cd4(mpda)4] loop. Adjacent loops are further connected
The crystal structure was solved by direct methods, using SHELXS-2014
through [Cd2(tib)2] motify to generate a 1D polymer chain (Fig. 2b).
and least-squares refined with SHELXL-2014 using anisotropic thermal
Seen from the crystal packing diagram of CP 2, the adjacent 1D poly­
displacement parameters for all non-hydrogen atoms [28,29]. Further
meric ribbon motif are bonded together by strong intermolecular forces
details for structural analysis are summarized in Table 1. Selected bond
to create a 3D supramolecular network (Fig. 2c).
distances and bond angles are listed in Table S1. Crystallographic data
(excluding structure factors) for the structures in this paper have been
3.1.3. Structure of [Cd(ppda) (tib)]⋅4H2O (3)
deposited with the Cambridge Crystallographic Data Centre, CCDC, 12
Single-crystal X-ray diffraction analysis reveals that CP 3 crystallizes
Union Road, Cambridge CB21EZ, UK. Copies of the data can be obtained
in the monoclinic system, P21/c space group, and the asymmetric unit of
free of charge on quoting the depository CCDC numbers 2132792,
CP 3 contains one Cd(II) atom, one fully deprotonated ppda2− ligand,
2132793, 2132794, for 1–3 (Fax: +44-1223-336-033; E-Mail: dep
one tib ligand and four lattice water molecules. As shown in Fig. 3a, each
[email protected], http://www.ccdc.cam.ac.uk).
Cd(II) atom in CP 3 is six-coordinated by three N-atom donors from three
tib ligands and three O atoms from two different ppda2− ligands. The tib
ligand acting as a three connected node joins the Cd(II) ions to form a 2D
extended structure. The ppda ligand in CP 3 shows a cis conformation

3
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

Fig. 1. (a) Coordination environment of Cd(II) atom in CP 1. All hydrogen atoms are omitted for clarity. (b) Cd(II) atoms are linked by tib ligands to form a 2D layer
in CP 1. (c) View of the 3D network of CP 1 extended by the 2D layers in an offset stacking fashion. The tib ligands in adjacent layers have been shown in green color,
the 1,2-pda ligands joining the layers have been shown in blue.

Fig. 2. (a) Coordination environments of Cd(II) atom in CP 2. All hydrogen atoms are omitted for clarity. (b) Ball-and-stick view of 1D loop-containing chain in CP 2.
(c) View of 3D extended supramolecular network in CP 2 (Layers show different color for clarity).

with two carboxyl groups exists on the same sides of the benzene ring ppda). The pda ligands display a range of conformations in the products.
plan and acts as a “V” type building block (Fig. S2c). Thus, the “V” type Both mpda and ppda ligands show cis-conformations (CPs 2, 3). In
building block link the Cd (II) atoms giving rise to a 1D wave chain, and contrast, opda shows trans-conformations (CP 1), this is due to the steric
these 1D chains run regularly through the 2D plane (Fig. 3b). The hindrance effect of phenyldiacetic acid. Three different coordination
adjacent 2D polymeric ribbon motifs are bound together by strong modes of phenyldiacetic acid isomers form different crystal structures
intermolecular forces to create a 3D supramolecular network. (3D, 1D, 2D for CPs 1, 2, 3 respectively). In this sense, the multi­
In total, three Cd(II) coordination polymers containing a 1,3,5-tris(1- carboxylate isomers play an important role in governing the coordina­
imidazolyl)benzene (tib) and a phenylenediacetate (pda) ligand have tion clusters and the final polymer structures in such assembled systems.
been prepared and structurally characterized. The compounds contain
one tib and one of three structural isomers of pda (opda, mpda and

4
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

Fig. 3. (a) Coordination environments of Cd(II) atom in CP 3. All hydrogen atoms are omitted for clarity. (b) Ball-and-stick view of 2D layer in CP 3.(Chains and
plane show different color for clarity). (c) View of 3D extended supramolecular network in CP 3 (Layers show different color for clarity).

3.2. Thermal analysis of CPs 1–3 close to the calculated 21.41%, 22.11%, 19.87%, based on CdO.

To examine the stability of CPs 1–3, thermal gravimetric analysis


(TGA) was performed. As shown in Fig. S3, the TGA curves exhibit a 3.3. Luminescent properties of CPs 1–3
weight loss of 3.21% (25–300 ◦ C) for CP 1, 5.14% (25–210 ◦ C) for CP 2,
2.82% (27–210 ◦ C) for CP 3 corresponding to the release of coordinated Since CPs 1–3 are constructed with a d10 electron configured Cd2+
water molecules and/or free water molecules (Calcd. 3.0%, 5.35% ion and fluorescent multi-carboxyl, multi-imidazole organic ligands
2.75% for CPs 1–3 respectively), followed by a sudden decrease in the containing aromatic or conjugated π moieties, it is highly possible for
weight. For CPs 1–3 the final residues of 22.13%, 23.10%, 20.17% are them to exhibit photoactivity. To examine the luminescent properties of
these d10 metal complexes, the luminescence spectra of CPs 1–3, the free

Fig. 4. (a) Solid-state emission spectra and (b) CIE-1931 chromaticity diagram of complexes 1–3 at room temperature.

5
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

opda, mpda, ppda and tib ligands have been measured at room tem­ nitrogen adsorption capacity. Barrett-joyner-halenda (BJH) was used to
perature. As shown in Fig. 4, the free opda, mpda and ppda ligands analyze the data and calculate the pore size.Total pore volume was
display main emission peak at 486 nm (λex = 300 nm), 501 nm (λex = measured at P/P0 = 0.99.
375 nm), and 479 nm (λex = 384 nm) respectively. While tib ligand
display main emission band at 364.8 nm (λex = 286 nm). In contrast, the
3.5. Sensing of TEC
emission peak of CP 1 [332 nm (λex = 280 nm)], CP 2 [333 nm (λex =
272 nm)], CP 3 [333 nm (λex = 283 nm)] are close to that of pda ligands,
Because of the excellent water stability and strong fluorescence of
and the waveform is similar. This indicates that the photoluminescence
CPs 1–3, they possible to be an ideal candidate for sensing material in
of CPs 1–3 could be mainly attributed to pda ligand centred electronic
water system. As shown in Fig. 5a and Fig. S7a, TEC has a conspicuous
excitation perturbed by the Cd2+ ion and other components. These shifts
quenching effect on the emission of CPs 1–3, whereas other seven an­
of the emission bands can be attributed to the deprotonated effect of pda
tibiotics display much smaller or minor influence on the emission.
and the coordination interactions of the organic ligands to Cd2+ ions as
Compared with CPs 1–3, CP 1 has higher quenching percentage, this
well as C–H⋅⋅⋅π interactions in CPs network [30,31]. The emission in­
may be attributed to the higher adsorption capacity of CPs with 3D
tensities of CPs 1–3 are much stronger than that of a free pda ligand, that
porous structure (18.140 m2/g). It is also necessary to carry out quan­
is the formation of complexes enhances the fluorescence of the ligand.
titative detection of the target object. The SV plot had a good linear
This phenomenon is a typical example of aggregation-induced emission
correlation at low concentrations, but diverges from linear upward
(AIE) and may be explained with that the coordination to metal ions in a
bending at high concentration (Fig. 5b and c and Figs. S7b and 7c). By
complexes restricted the deformation of the ligand and the induced
monitoring the luminescent intensities for CPs 1–3, the values of Ksv, R2,
nonradiative relaxation [32–34].
and LOD are determined and listed in Table 2. According to Table 3, the
sensitivity in this work is superior/near to most of the reported fluo­
3.4. BET analysis of CP 1 rescence probes for TEC and meets the standard of World Health Or­
ganization (WHO) and occupational safety and health administration
The surface area, pore size, pore volume and other parameters of the
material were analyzed by using N2 adsorption-desorption and non-local Table 2
density functional theory (NLDFT) methods. It can be seen from Fig. S5 KSV, R2, and LOD of CPs 1–3 toward TEC at room temperature.
that the nitrogen adsorption-desorption isotherm of CP 1 is a typical H3 antibiotic TEC
hysteresis type IV isotherm, which indicates the existence of mesopores CP 1 2 3
and macropores in the material [35]. CP 1 has a large specific surface 5
KSV(10 L/mol) 3.059 3.248 3.852
(18.14 m2 g− 1) area and pore volume (0.041 cm3 g− 1), which provides R2 0.9950 0.9975 0.9927
good conditions for the adsorption of tetracycline. At the same time, the LOD( × 10− 7 M) 2.1 3.19 5.6
larger pores (7.61 nm) also facilitate the entry of tetracycline into the CP LOD(ppb) 93.2 141.6 248.6
1. Brunauer - Emmett - Teller (BET) method was used to determine

Fig. 5. (a) Luminous intensity of CP 1 (2 mL, prepared solution) in aqueous solutions with different antibiotic (0.001 M). (b) Concentration-dependent luminescence
quenching of CP 1 suspension in water by TEC (excited at 257 nm). (c) Stern− Volmer plot of CP 1 suspension quenched by TEC.

6
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

Table 3
Comparison of luminescent materials for sensing TEC.
materials objectives KSV LOD (ppb) refs
5
[Zn3(L)2(1,4-bimb)3]n TEC 1.98 × 10 0.15 μМ [36]
In-sbdc TEC 3.11 × 105 0.28 μМ [37]
PCN–128Y TEC 9.84 × 105 0.03 μМ (30 nM) [38]

(OSHA) for the maximum allowable level of antibiotics in drinking


water [39]. Furthermore, CPs 1–3 have high sensitivity, selectivity, and
resistance to interference from other coexisting antibiotics in detecting
TEC in aqueous solution (Fig. 6a, Fig. S8a). And the fluorescence
quenching progress is very fast (Fig. S9), the present fluorescent probe
exhibited good cycle stability. After 5 cycles, the quenching efficiency is
decreased very little (Fig. 6b, Fig. S8b). These results indicate that CPs
1–3 can serve as sensitive sensor for the selective and rapid detection of
TEC in water system.

3.6. Fluorescence quenching mechanism to TEC


Fig. 7. The UV absorption spectra of various antibiotics (0.1 mM) and the
emission spectrum of CPs 1–3 in aqueous solution.
Usually, the photoinduced electron transfer (PET) and the inner filter
effect (IFE) can cause the fluorescence quenching behavior of the CPs
fluorescent materials. overlaps with the emission peak of CPs 1–3. This indicates that TEC can
When the fluorescent material bears the higher LUMO energy level strongly absorb the emission light of CPs 1–3. As a result, the lumines­
than the analyte, the electron transfer from the fluorescent material to cence of CPs 1–3 is highly sensitively quenched. The above facts imply
the analyte occurs, which is considered as the PET effect causing the that the inner filter effect (IFE) exists between CPs 1–3 and TEC.
fluorescence quenching [40]. Fig. S10 lists the HOMO and LUMO energy Compared with CPs 2, 3, the emission spectrum of CP 1 has higher
levels of various antibiotics and CPs 1–3 calculated according to the overlap with the UV–Vis absorption of TEC, this may be attributed to the
density functional theory (DFT) method [41]. Compared with other higher adsorption capacity of CP 1 with 3D porous structure. Obviously,
antibiotics, the LUMO energy level of CEF, CHL, PEN and TEC is lower the photoinduced electron transfer (PET) and the inner filter effect (IFE)
than that of CPs 1–3. The electron transfer between complex (donor) and simultaneously exist between CPs 1–3 and TEC, which leads to high
CEF, CHL, PEN and TEC (acceptor) happens, which prevents the exci­ sensibility and selectivity for detection of TEC.
tation electron back to the ground state. This leads to different fluo­ As the fluorescence lifetime of CPs 1–3 remained unchanged with the
rescence quenching effects. However, the quenching order of the increase of TEC concentration, the sensing process should be a static
fluorescence is not in accordance with the LUMO energy levels of anti­ quenching process rather than a dynamic quenching process, which is
biotics. Especially, CHL has the lowest LUMO energy level, but does not consistent with literature reports [47]. (Table S3).
display the highest fluorescence quenching effect, implying that the PET
effect is not unique reason causing the fluorescence quenching in the 3.7. Adsorption of tetracycline(TEC)
current system.
When the excitation and/or emission lights of the fluorescent matter 3.7.1. Adsorption kinetics study
are absorbed by the analyte, the fluorescence IFE happens [42–46]. Fig. 8 shows the change of TEC concentration over time. The
Fig. 7 displays the absorption spectra of tested antibiotics and the adsorption capacity increases with the increase of the contact time.
emission spectrum of CPs 1–3 in aqueous solution. It is clearly found that When the adsorption time reaches 1440 min, the adsorption capacity is
an absorption peak of TEC located in the region of 275–320 nm well saturated. In the subsequent adsorption process, the adsorption capacity

Fig. 6. Luminescence quenching efficiency of CP 1 with 1 mM antibiotic aqueous solution upon adding different concentrations of TEC (a). The cycle stability of CP 1
fluorescent probe for the detection of TEC (b).

7
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

Table 4
Kinetic parameters for TCE adsorption onto CP 1.
Kinetic Models Parameters

pseudo-first-order qe,exp (mg qe,cal (mg k1 R2


kinetic g− 1) g− 1) (min− 1)
64.493 32.128 0.00181 0.74076
pseudo-second-order qe,cal (mg k2 (min− 1) R2
kinetic g− 1)
60.001 0.000203 0.98668
− 1 − 0.5 − 1
Intraparticle diffusion model Kid (mg g min ) C (mg g ) R2

Step 1 3.9557 1.57679 0.85539


Step 2 0.8508 30.00299 0.99315
Step 3 0.92153 26.95368 0.99244

experimental value of qe,exp (64.493 mg g− 1). The adsorption data of CP


1 has a better fit with the pseudo-second model, this result shows that
chemical adsorption dominates the adsorption process of Cp 1 for
tetracycline [50].
The internal diffusion model can be used to analyze the control steps
in the adsorption process. There are three linear parts in Fig. 9c. The first
Fig. 8. The effect of materials on the adsorption of tetracycline under different linear part illustrates the transfer and diffusion behavior of tetracycline
time. Experimental conditions: The amount of adsorbent added is 0.25 g L− 1 on the adsorbent surface. The second linear part is the process of
(25 ◦ C, initial pH = 7). tetracycline diffusion from the outer surface of Cp 1 to its internal pores
and capillary structure. From the slope of the line segment, it can be seen
does not change significantly. The adsorption capacity of CP 1 is 64.493 that the adsorption rate is significantly reduced and is the lowest among
mg g− 1, which is similar to the adsorption capacity of MIL-53(Fe) for the three sets of line segments. Therefore, this process is determined by
tetracycline [48]. the rate of the adsorption process step. The third linear part shows that
Adsorption kinetics of CP 1 for tetracycline can be described by tetracycline occupies the adsorption site of the adsorbent and gradually
pseudo-first-order, pseudo-second-order, and intra-particle diffusion reaches the adsorption saturation state. In addition, none of the three
models, respectively. The adsorption data was fitted with three typical line segments passes through the origin coordinates, indicating that
kinetic equations to study the adsorption rate, adsorption type and rate intra-particle diffusion is not the only step to control this adsorption
control steps of the entire adsorption process. The nonlinear equations of process [51].
the three dynamics are described by eqs (2)–(4) respectively [49].
3.7.2. CP 1 adsorption isotherm for TEC
pseudo-first-order kinetic: ln(qe - qt) - k1t (2) Adsorption of tetracycline with different initial concentrations
(5–200 mg/L) were investigated at room temperature (pH = 7.0). With
pseudo-second-order kinetic: t/qt = 1/k2qe2 + t/qe (3)
the increase of the initial tetracycline concentration, the adsorption
intraparticle diffusion model: qt = kidt 1/2
+C (4) capacity of CP1 has been significantly improved (Fig. 10). This is
because a higher initial concentration can increase a stronger adsorption
Where qe is the equilibrium adsorption capacity of CPs, qt (mg g− 1) is the driving force [52].
adsorption capacity at t (min), and C is the model constant of intra - In order to study the equilibrium relationship between the remaining
particle diffusion. k1 (min− 1), k2 (g min− 1 mg− 1) and kid (mg g− 1 min− 1/ TEC concentration (Ce) and the adsorption amount (qe) of tTEC, three
2
) are the rate constants of the quasi-first-order kinetic model, the quasi- isotherm models Langmuir, Freundlich and Temkin were used to fit the
second-order kinetic model and the intra particle diffusion model, experimental data, respectively [53].
respectively.
The fit of the three models is shown in Fig. 9. The kinetic parameters Langmuir:qe = qmkLCe/(1+kLCe) (5)
of the three models are also sorted out in Table 4. The R2 value of the
Freundlich:qe = kFCe1/n (6)
pseudo-first-level model (0.7407) is much smaller than that of the
pseudo-second-level model (0.9866), and the calculated qe,cal (60.001 Temkin:qe = RTln(kTCe)/b (7)
mg g− 1) of the pseudo-second-order kinetic equation is closer to the

Fig. 9. Pseudo-first-order kinetic model (a), pseudo-pseudo-second-order kinetic model (b), and intra-particle diffusion model (c) for tetracycline adsorption.
Experimental conditions: the amount of adsorbent added is 0.25 g L− 1 (25 ◦ C, initial pH = 7).

8
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

Table 5
Langmuir, Freundlich and Temkin isotherms parameters for TEC adsorption
onto CP 1.
Model Parameters

Langmuir isotherm qm,exp qm,cal KL RL R2


177.002 182.672 0.00501 0.83 0.98892
Freundlich isotherm N KF R2
1.427 4.61286 0.9759
Temkin isotherm B KT R2
52.1236 0.1373 0.8958

where C0 (mg g− 1) is the initial TEC concentration. The value of this


parameter is beneficial to the adsorption process when 0<RL < 1, but the
adsorption process becomes unfavorable when RL > 1 or RL < 0. The
value of RL in this experiment is 0.83, indicating that the adsorption
behavior was favorable.

3.7.3. CP 1 adsorption thermodynamics for TEC


Temperature is a key factor affecting adsorption. In order to study
the effect of temperature on adsorption rate and equilibrium adsorption
Fig. 10. Adsorption performance of CP 1 at different initial concentrations. capacity, we conducted adsorption behavior at 25, 35, and 45 ◦ C
Experimental conditions: The amount of adsorbent added is 0.25 g L− 1(25 ◦ C, (Fig. S11). With the increase of temperature, the equilibrium adsorption
initial pH = 7). capacity and adsorption rate of CP 1 for TEC have been improved. The
Gibbs free energy (ΔG), enthalpy change (ΔH) and entropy change (ΔS)
Where qe is consistent with the constant in equation, Ce (mg L− 1) is the can be calculated according to the following eqs (9)–(12) [56–58].
equilibrium concentration, KL (L mg− 1) and qm (mg g− 1) are Langmuir
ΔG = -RTlnkα (9)
constant and Langmuir equilibrium adsorption capacity respectively. KF
[(L mg− 1)1/n mg g− 1] and n are Freundlich parameters, which are Kα = 1 0 kL6
(10)
related to adsorption capacity and strength, respectively. KT (L mg-1) is
Temkin isotherm constant, and b (J mol− 1g mg− 1) is adsorption heat lnCe = ΔH/RT + kα (11)
constant. R (8.314 J mol− 1 K− 1) is the gas constant and T (K) is the
ΔS = (ΔH - ΔG)/T (12)
absolute temperature.
The Langmiur model assumes that the adsorption sites on the − 1 − 1
Where R (8.314 J mol K ) is the gas constant, T (K) is the thermo­
adsorbent surface are uniformly distributed, and there is no interaction dynamic temperature, Ce (mg L− 1) is the equilibrium concentration, and
between the adsorbates. It is a single-layer adsorption that occurs on the kα is the thermodynamic equilibrium constant without unit, which can
outer surface of the adsorbent [54]. The three adsorption isotherm be obtained by multiplying the Langmiur equilibrium constant kL by 106
models are shown in Fig. 11, and the adsorption isotherm parameters of [59,60].
CP 1 are shown in Table 5. It can be seen from the isotherm parameters The thermodynamic parameter values are shown in Table 6. The
that the nonlinear correlation coefficient (R2) of the Langmiur isotherm standard Gibbs free energy (ΔG) is less than zero at different
is 0.98892, which fits well with the experimental data. This shows that
the adsorption of tetracycline is a monolayer adsorption on the uni­
formly distributed active sites on the surface of CP 1. Table 6
Thermodynamic parameters for TEC adsorption onto CP 1.
In addition, the dimensionless constant RL is used to express the basic
characteristics of the Langmiur isotherm. RL is called the equilibrium ΔG (KJ mol− 1) ΔH (KJ mol− 1) 1
Analyte T (K) ΔS (J mol− K− 1 )
parameter or separation factor [55], which can be expressed in eq (8). TEC 298 − 21.107 10.291 0.105
308 − 21.815
RL = 1/(1 + kLC0) (8) 318 − 22.523

1
Fig. 11. Langmuir model (a), Freundlich model (b) and Temkin model (c) of TEC adsorption. Experimental conditions: The amount of adsorbent added is 0.25 g L−
(25 ◦ C, initial pH = 7).

9
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

experimental temperatures, indicating that the adsorption process of CP


1 for tetracycline is spontaneous. With the temperature increasing, the
Gibbs free energy (ΔG) decreases, which is beneficial to the adsorption
process. The value of ΔH is positive, suggesting that the adsorption re­
actions were endothermic.The value of ΔS is positive, indicating that the
randomness of the interface between CP 1 and TEC increases during the
adsorption process. In summary, TEC adsorption onto CP 1 is a spon­
taneous endothermic process.

3.7.4. The effect of pH


The ionization degree of TEC and the surface charge of CP 1 are
easily affected by the solution pH, which will interfere with the
adsorption process [61]. The adsorption capacity of CP 1 for tetracycline
at different pH (3.0–9.0) is shown in Fig. 12. CP 1 has the maximum
adsorption capacity (62.1586 mg/L) for TEC at pH = 7. The tetracycline
PKa are 3.32, 7.78 and 9.58, respectively, which means that tetracycline
is positively charged when pH < 3.32, negatively charged when pH >
7.78, and is neutral between 3.32 < pH < 7.78. According to the zeta
potential of CP 1 at different pH, the point of zero charge (PZC) of CP 1 is
about 7.1. When the pH is higher than PZC 7.1, the surface of the
adsorbent is positively charged, otherwise it is negatively charged.
Fig. 13. Effect of salt content on TEC removal. Experimental conditions: The
Therefore, there is no electrostatic interaction between CP 1 and TEC at
amount of adsorbent added is 0.25 g L− 1 (25 ◦ C, initial pH = 7).
pH < 7.78, and the adsorption capacity decreases due to electrostatic
repulsion when pH > 7.78. CP 1 still has high adsorption capacity for
3.7.7. CP 1 adsorption mechanism
TEC under different pH, which shows that electrostatic interaction is not
The structural properties of CP 1 and the interaction with TEC are
the main adsorption mechanism.
considered to be the main mechanism of adsorption. The pore size of CP
1 is larger than the molecular diameter of TEC (0.5–2 nm) [62], which
3.7.5. The effect of ionic strength
makes it easy for tetracycline to enter CP 1 and bind to the adsorption
The effect of ionic strength on the adsorption process was investi­
site. The experimental data of CP 1 adsorption of TEC fits better with
gated. Fig. 13 summarizes the TEC adsorption over CP 1 in the presence
pseudo-second-order kinetics, indicating that it is mainly chemical
of NaCl at different concentrations. As the concentration of NaCl
adsorption. The main effects of chemical adsorption include electro­
increasing, the adsorption capacity decreases. Sodium ions compete
static interactions, hydrogen bonds, and π-π interactions [63]. It can be
with TEC molecules for active sites on CP 1 surface, resulting in elec­
seen from the adsorption of tetracycline by CP 1 at different pH that the
trostatic shielding. NaCl can also increase the solution viscosity and
adsorption capacity is reduced due to the electrostatic repulsion at pH >
density, impeding the transfer of TEC molecules to CP 1.
7, but the tetracycline can still be adsorbed effectively, which shows that
electrostatic interaction is not the main mechanism of the adsorption
3.7.6. Reusability of CP 1
process. The π-π interaction between CP 1 and the benzene ring of TEC is
Reusability is an important indicator for judging whether the
considered to be the main adsorption mechanism.
adsorbent has application value. As shown in Fig. S12, after three cycles,
CP 1 still maintains about 80% of the initial adsorption capacity, which
has good recyclability. The decrease in adsorption capacity is due to the 4. Conclusion
strong occupation of the surface active site of CP 1 by uneluted TEC
molecules. The loss of material in the process of adsorption and In summary, CPs 1–3 have been synthesized successfully under hy­
desorption is also the reason for the decrease of CP 1 adsorption drothermal conditions by the reaction of Cd(II) salt and 1,3,5-tris(1-imi­
capacity. dazolyl)benzene in the presence of different flexible phenylenediacetate
isomers. In CP 1, the opda2− ligand exhibits trans-conformation and act
as bridging building block extending 2D (Cd-tib) layers into a 3D
network. While mpda2− and ppda2− ligands in CP 2 and CP 3 possess cis-

Fig. 12. (a). Effect of solution pH on TEC adsorption by CP1. (b). ZETA potential of CP 1 under different pH. Experimental conditions: The amount of adsorbent
added is 0.25 g L− 1 (25 ◦ C, initial pH = 7).

10
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

conformation and can be seen as “V” type building block, linking Cd(II) [12] Husnain SM, Asim U, Yaqub A, Shahzad F, Abbas N. Recent trends of MnO2-derived
adsorbents for water treatment: a review. New J Chem 2020;44:6096–120.
atoms to form 1D and 2D structure with tib ligand. The diversities of the
[13] Javad I, Mohsen M, Mohammad D, Mohammad RE. Adsorption and desorption of
structures for CPs 1–3 have some effects on fluorescence sensing prop­ amoxicillin antibiotic from water matrices using an effective and recyclable MIL-53
erty. The results reveal that CP 1 possesses higher quenching percentage (Al) metal− organic framework adsorbent. J Chem Eng Data 2021;66:389–403.
toward tetracycline(TEC), because CP 1 possesses 3D porous structure. [14] Seyed MM, Payam S, Sedigheh Z, Mohammad RR. Tetracycline antibiotic removal
from aqueous solutions by MOF-5: adsorption isotherm, kinetic and
Furthermore, the adsorption performance of CP 1 to TEC was studied. thermodynamic studies. J Environ Chem Eng 2018;6:6118–30.
The maximum adsorption capacity was 64.493 mg g− 1 (PH = 7.0, T = [15] Peng X, Hu F, Zhang T, Qiu F, Dai H. Aminefunctionalized magnetic bamboo-based
298K). The adsorption isotherms were described accurately by the activated carbon adsorptive removal of ciprofloxacin and norfloxacin: a batch and
fixed-bed column study. Bioresour Technol 2018;249:924–34.
Langmuir model, indicating a monolayer adsorption process. And the [16] Yang W, Lu Y, Zheng F, Xue X, Li N, Liu D. Adsorption behavior and mechanisms of
adsorption is driven by intraparticle diffusion. This study expands the norfloxacin onto porous resins and carbon nanotube. Chem Eng J 2012;179:112–8.
number 3D CPs available for restoring the safety of water environments. [17 ] Wang B, Jiang YS, Li FY, Yang DY. Preparation of biochar by simultaneous
carbonization, magnetization and activation for norfloxacin removal in water.
Bioresour Technol 2017;233:159–65.
CRediT authorship contribution statement [18] Liang Z, Zhao Z, Sun T, Shi W, Cui F. Adsorption of quinolone antibiotics in
spherical mesoporous silica: effects of the retained template and its alkyl chain
length. J Hazard Mater 2016;305:8–14.
Mei-Li Zhang: Experimental design, Synthesis, Characterization, [19] Wu X, Huang M, Zhou T, Mao J. Recognizing removal of norfloxacin by novel
Formal analysis, Writing – review & editing. Xue-Ying Lu: Experimental magnetic molecular imprinted chitosan/γ-Fe2O3 composites: selective adsorption
design, Formal analysis, Writing – original draft. Ye Bai: Fluorescence mechanisms, practical application and regeneration. Separ Purif Technol 2016;
165:92–100.
and adsorption experiments, Formal analysis. Yi-Xia Ren: Fluorescence
[20] Dutta A, Pan Y, Liu JQ, Kumar A. Multicomponent isoreticular metal-organic
mechanism, Adsorption model. Ji-Jiang Wang: Fluorescence mecha­ frameworks: principles, current status and challenges. Coord Chem Rev 2021;445:
nism, Adsorption model. Xiao-Gang Yang: Writing – review & editing. 214074.
[21] Liu JQ, Luo ZD, Pan Y, Kumar A. Trivedi M. Recent developments in luminescent
coordination polymers: designing strategies, sensing application and theoretical
Declaration of competing interest evidences. Coord Chem Rev 2020;406:213245.
[22] Qin ZS, Dong WW, Zhao J, Wu YP, Zhang Q, Li DS. A water-stable Tb(III)-based
metal-organic gel (MOG) for detection of antibiotics and explosives. Inorg Chem
The authors declare that they have no known competing financial Front 2018;5:120.
interests or personal relationships that could have appeared to influence [23] Zhou SH, Lu L, Liu D, Wang J, Sakiyama H, Muddassir M, Nezamzadeh-Ejhieh A,
the work reported in this paper. Liu JQ. Series of highly stable Cd(ii)-based MOFs as sensitive and selective sensors
for detection of nitrofuran antibiotic. CrystEngComm 2021;23:8043–52.
[24] Zhong YY, Chen C, Liu S, Lu CY, Liu D, Pan Y, Sakiyama H, Muddassir M, Liu JQ.
Acknowledgment A new magnetic adsorbent of eggshell-zeolitic imidazolate framework for highly
efficient removal of norfloxacin. Dalton Trans 2021;50:18016–26.
[25] Azarpira H, Mahdavi Y, Khaleghi O, Balarak D. Thermodynamic studies on the
This work was financially supported by the National Natural Science
removal of metronidazole antibiotic by multi-walled carbon nanotubes. Der Pharm
Foundation of China (21573189 and 21971100), the Nature Scientific Lett 2016;8:114− 121.
Foundation of the Shaanxi Province of China (2016XT-24), and the [26] Reza A, Ali RM, Jafar S. Synthesis of a nanostructured pillar MOF with high
Natural Scientific Research Foundation of National Undergraduate adsorption capacity towards antibiotics pollutants from aqueous solution. J Hazard
Mater 2019;366:439–51.
Training Programs of Innovation and Entrepreneurship [27] Jia XN, Li SJ, Wang YD, Wang T, Hou XH. Adsorption behavior and mechanism of
(S201910719058). sulfonamide antibiotics in aqueous solution on a novel MIL-101(Cr)@GO
composite. J Chem Eng Data 2019;64:1265–74.
[28] Sheldrick G. Program for crystal structure Determination. 2014. SHELXL-2014/7.
Appendix A. Supplementary data [29] Sheldrick G. Program for crystal structure Refinement. 2014. SHELXL-2014/7.
[30] Yang XG, Zhai ZM, Lu XM, Qin JH, Chang XH, Han ML, Li FF, Ma LF. π-Type
Supplementary data to this article can be found online at https://doi. halogen bonding enhanced long-last room temperature phosphorescence of Zn(II)
coordination polymers for photoelectron response applications. Inorg Chem Front
org/10.1016/j.dyepig.2022.110172. 2020;7:2224–30.
[31] Yang XG, Zhai ZM, Lu XM, Qin JH, Li FF, Ma LF. Hexanuclear Zn(II) induced dense
References π-stacking in MOF featuring long-last room temperature phosphorescence. Inorg
Chem 2020;59:10395–9.
[32] Hong W, Li L, Xue R, Xu X, Wang H, Zhou J, Zhao H, Song Y, Liu Y, Gao J. One-pot
[1] Cui Y, Kang W, Qin L, Ma J, Liu X, Yang Y. Ultrafast synthesis of magnetic hollow
hydrothermal synthesis of Zinc ferrite/reduced graphene oxide as an efficient
carbon nanospheres for the adsorption of quinoline from coking wastewater. New J
electrocatalyst for oxygen reduction reaction. J Colloid Interface Sci 2017;485:
Chem 2020;44:7490–500.
175–82.
[2] Balarak D, Azarpira H. Photocatalytic degradation of Sulfamethoxazole in water:
[33] Meisner J, Karwounopoulos J, Walther P, Kastne J, Naumann S. The lewis pair
investigation of the effect of operational parameters. Int J Chem Res 2016;9:731–8.
polymerization of lactones using metal halides and N-heterocyclic olefins:
[3] Balarak D, Mostafapour FK, Azarpira H. Adsorption isotherm studies of tetracycline
theoretical insights. Molecules 2018;23:432–46.
antibiotics from aqueous solutions by maize stalks as a cheap biosorbent. Int J
[34] Yang XG, Zhai ZM, Lu XM, Qin JH, Li FF, Ma LF. Hexanuclear Zn(II) induced dense
Pharm Technol 2016;8:16664–75.
π-stacking in MOF featuring long-last room temperature phosphorescence. Inorg
[4] Ahmadi S, Banach A, Mostafapour FK, Balarak D. Study urvey of cupric oxide
Chem 2020;59:10395–9.
nanoparticles in removal efficiency of ciprofloxacin antibiotic from aqueous
[35] Song HQ, Zhu Q, Zheng XJ. One-step synthesis of three-dimensional graphene/
solution: adsorption isotherm study. Desalination Water Treat 2017;89:297–303.
multiwalled carbon nanotubes/Pd composite hydrogels: an efficient recyclable
[5] Azarpira H, Balarak D. Rice husk as a biosorbent for antibiotic Metronidazole
catalyst for Suzuki coupling reactions. J Mater Chem 2015;19:10368–77.
removal: isotherm studies and model validation. Int J Chem Res 2016;9:566–73.
[36] Fan C, Zhang X, Li N, Xu C, Wu R, Zhu B, Zhang G, Bi S, Fan Y. Zn-MOFs based
[6] Pill WS, Nazmul AK, Sung HJ. Removal of nitroimidazole antibiotics from water by
luminescent sensors for selective and highly sensitive detection of Fe3+ and
adsorption over metal–organic frameworks modified with urea or melamine. Chem
tetracycline antibiotic. J Pharm Biomed Anal 2020;188:113444.
Eng J 2017;315:92–100.
[37] Liu Q, Ning D, Li WJ, Du XM, Wang Q, Li Y, Ruan WJ. Metal–organic framework-
[7] Beltran FJ, Aguinaco A, García-Araya JF, Oropesa A. Ozoné and photocatalytic
based fluorescent sensing of tetracycline-type antibiotics applicable to
processes to remove the antibiotic sulfamethoxazole from water. Water Res 2008;
environmental and food analysis. Analyst 2019;144:1916–22.
42:3799–808.
[38] Zhou Y, Yang Q, Zhang D, Gan N, Li Q, Cuan J. Detection and removal of antibiotic
[8] Adams C, Asce M, Wang Y, Loftin K, Meye M. Removal of antibiotics from surface
tetracycline in water with a highly stable luminescent MOF. Sensor Actuator B
and distilled water in conventional water treatment processes. J Environ Eng 2002;
Chem 2018;262:137–43.
128:253–60.
[39] Xu XJ, Yan B. Eu(III)-functionalized ZnO@MOF heterostructures: integration of
[9] Choi KJ, Son HJ, Kim SH. Ionic treatment for removal of sulfonamide and
pre-concentration and efficient charge transfer for the fabrication of a ppb-level
tetracycline classes of antibiotic. Sci Total Environ 2007;387:247–56.
sensing platform for volatile aldehyde gases in vehicles. J Mater Chem 2017;5:
[10] Ternes AT, Meisenheimer M, Mcdowell D, Frank S, Brauch H, Haistgulde B,
2215–23.
Preuss G, Uwe WA, Zuleiseibert N. Removal of pharmaceuticals during drinking
[40] Mallick A, Garai B, Addicoat MA, Petkov PS, Heine T, Banerjee R. Solid state
water treatment. Environ Sci Technol 2002;36:3855–63.
organic amine detection in a photochromic porous metal organic framework. Chem
[11] Sarker M, Bhadra BN, Seo PW, Jhung SH. Adsorption of benzotriazole and
Sci 2015;6:1420–5.
benzimidazole from water over a Co-based metal azolate framework MAF-5 (Co).
J Hazard Mater 2017;324:131–8.

11
M.-L. Zhang et al. Dyes and Pigments 202 (2022) 110172

[41] Frisch M, Trucks G, Schlegel H. Gaussian 09, Revision E.01. Wallingford CT: adsorption and visible-light-driven photocatalysis. Appl Catal B Environ 2016;186:
Gaussian, Inc.; 2013. 19–29.
[42] Lu BB, Jiang W, Yang J, Liu YY, Ma JF. Resorcin[4]arene-based microporous [52] Abdi J, Vossoughi M, Mahmoodi NM. Synthesis of metal-organic framework hybrid
metal-organic framework as an efficient catalyst for CO2 cycloaddition with nanocomposites based on GO and CNT with high adsorption capacity for dye
epoxides and highly selective luminescent sensing of Cr2O72-. ACS Appl Mater removal. Chem Eng J 2017;326:1145–58.
Interfaces 2017;9:39441–9. [53] Minisy I, Ayad MM, Salahuddin N. Chitosan/polyaniline hybrid for the removal of
[43] Gao DM, Wang ZY, Liu BB, Ni L, Wu MH, Zhang ZP. Resonance energy transfer- cationic and anionic dyes from aqueous solutions. J Appl Polym Sci 2018;136:
amplifying fluorescence quenching at the surface of silica nanoparticles toward 47056.
ultrasensitive detection of TNT. Anal Chem 2008;80:8545–53. [54] Chatterjee S, Lee DS, Lee MW, Woo SH. Enhanced adsorption of Congo red from
[44] Jin SS, Han X, Yang J, Zhang HM, Liu XL, Ma JF. Luminescent coordination aqueous solutions by chitosan hydrogel beads impregnated with cetyl trimethyl
polymers based on a new resorcin[4]arene-functionalized tetracarboxylate: highly ammonium bromide. Bioresour Technol 2009;10:2803–9.
selective luminescent detection of metal cations, anions and small organic [55] Weber TW, Chakravorti RK. Pore and solid diffusion models for fixed-bed
molecules. J Lumin 2017;188:346–55. adsorbents. AIChE J 1974;20:228–38.
[45] Lin ZJ, Zheng HQ, Zheng HY, Lin LP, Xin Q, Cao R. Efficient capture and effective [56] Liu Y, Liu YJ. Biosorption isotherms, kinetics and thermodynamics. Separ Purif
sensing of Cr2O72- from water using a zirconium metal-organic framework. Inorg Technol 2008;61:229–42.
Chem 2017;56:14178–88. [57] Liu Y, Xu H. Equilibrium, thermodynamics and mechanisms of Ni2+ biosorption by
[46] Rachuri Y, Parmar B, Bisht KK, Suresh E. Multiresponsive adenine-based aerobic granules. Biochem Eng J 2007;35:174–82.
luminescent Zn(II) coordination polymer for detection of Hg2+ and trinitrophenol [58] Yu L. Is the free energy change of adsorption correctly calculated? J Chem Eng
in aqueous media. Cryst Growth Des 2017;17:1363–72. Data 2009;54:1981–5.
[47] Buragohain A, Yousufuddin M, Sarma M, Biswas S. 3D luminescent [59] Milonjic SK. A consideration of the correct calculation of thermodynamic
amidefunctionalized cadmium tetrazolate framework for selective detection of parameters of adsorption. J Serb Chem Soc 2007;72:1363–7.
2,4,6- trinitrophenol. Cryst Growth Des 2016;16:842–51. [60] Jin JH, Yang ZH, Xiong WP. Cu and Co nanoparticles co-doped MIL-101 as a novel
[48] Xiong WP, Zeng GM, Yang ZH, Zhou YY, Zhang C. Adsorption of tetracycline adsorbent for efficient removal of tetracycline from aqueous solutions. Sci Total
antibiotics from aqueous solutions on nanocomposite multi-walled carbon Environ 2019;650:408–18.
nanotube functionalized MIL-53 (Fe) as new adsorbent. Sci Total Environ 2018; [61] Parolo ME, Savini MC, Vallés JM. Tetracycline adsorption on montmorillonite: pH
627:235–44. and ionic strength effects. Appl Clay Sci 2008;40:179–86.
[49] Nie Y, Hu C, Kon C. Enhanced fluoride adsorption using Al(III) modified calcium [62] Helio F, Dos S, Wagner BDA, Michael CZ. Conformational analysis of the
hydroxyapatite. J Hazard Mater 2012;10:194–9. anhydrotetracycline molecule: a toxic decomposition product of tetracycline.
[50] Zhao X, Liu D, Huang H. The stability and defluoridation performance of MOFs in J Pharmaceut Sci 1998;87:190–5.
fluoride solutions. Microporous Mesoporous Mater 2014;185:72–8. [63] Oveisi M, Asli MA, Mahmoodi NM. MIL-Ti metal-organic frameworks (MOFs)
[51] Wang H, Yuan X, Wu Y. In situ synthesis of In2S3@MIL-125(Ti) core-shell nanomaterials as superior adsorbents: synthesis and ultrasound-aided dye
microparticle for the removal of tetracycline from wastewater by integrated adsorption from multicomponent wastewater systems. J Hazard Mater 2017;347:
123–40.

12

You might also like