Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Membrane Science 679 (2023) 121706

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Smart integration of MOFs and CQDs to fabricate defect-free and


self-cleaning TFN membranes for dye removal
Die Ling Zhao a, Haiyi Jin a, Qipeng Zhao b, Yanchao Xu a, Liguo Shen a, Hongjun Lin a, **,
Tai-Shung Chung b, c, *
a
College of Geography and Environmental Sciences, Zhejiang Normal University, Jinhua, 321004, China
b
Department of Chemical and Biomolecular Engineering, National University of Singapore, 117585, Singapore
c
Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, 10607, Taiwan

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

Keywords: The surface chemistry and functionality of metal-organic frameworks (MOFs) can be manipulated through a
Metal-organic frameworks judicious choice of integration with other nanomaterials. In this work, heterojunctions of UiO-66-NH2 and
Carbon quantum dots carbon quantum dots (CQDs) were prepared to explore their synergetic advantages as nanofillers in thin-film
Thin-film nanocomposite
nanocomposite (TFN) membranes for dye removal. The successful integration of UiO-66-NH2/CQD were veri­
Nanofiltration
fied by TEM, FT-IR and XPS. The uniform loading of ultra-small CQDs on the surfaces of UiO-66-NH2 can not only
Self-cleaning
Dye removal enhance their compatibility with the polyamide matrix, but also promote their photocatalytic oxidation under
visible light. Therefore, comparing to the pristine thin-film composite membranes, the optimal TFN membranes
embedded with 0.2 wt% of UiO-66-NH2/CQD have demonstrated remarkable improvements in pure water
permeance by 89.7% and higher selectivity to dyes of different charges and molecular weights. In addition, the
decrease of water flux caused by dye adsorption during the long-term filtration process can be recovered by
92–99% via photocatalytic degradation of the attached dyes within 10 min. This work may offer a new strategy
for fabricating defect-free and self-cleaning nanofiltration membranes for dye removal and wastewater
reclamation.

1. Introduction accumulate, yielding a declined membrane permeability and separation


efficiency [12,13]. Therefore, it is imperative to develop NF membranes
Treatment and recycle of industrial wastewater are of great impor­ with both high performance and self-cleaning ability for dye wastewater
tance given their severe menace to the environment and the growing treatment [14–18].
global water scarcity [1–5]. Among them, dye wastewater produced Metal-organic frameworks (MOFs) were intensively studied as
from textile, dye production and printing industries has raised people’s nanofillers in the selective layers [19–22]. With well-defined and highly
attention, while conventional treating methods (e.g., adsorption, tunable nano-sized channels, MOFs are able to sieve solutes larger than
advanced oxidation or microbial degradation) still face challenges such the channel but allow molecules with smaller sizes to traverse, rendering
as low efficiency and poor feasibility in dye recycle [6,7]. In recent them great feasibility of applications in the fields of gas and liquid
years, nanofiltration (NF), which can effectively repel dye molecules separation [23–26]. Additionally, benefitting from their relatively rigid
with relatively low molecular weights of <2000 Da, has emerged as a crystalline nature, MOFs can surpass organic porous species, such as
competitive and desirable alternative due to its ability to remove and covalent-organic frameworks (COFs) and polymers of intrinsic micro­
recycle the dyes simultaneously [8–10]. Nevertheless, the current porosity (PIMs), which are plagued by aging and reduction of porosity
dominant membranes for NF processes; namely, thin-film composite [27]. On the other hand, inorganic porous nanomaterials like zeolites
(TFC) membranes are still limited by the “trade-off” relationship and are suffering from less diversity in chemistry and structure, as well as
membrane fouling [11]. Especially in dye wastewater treatment, dye lower processability [28]. Therefore, featuring with relatively robust
molecules can anchor to the membrane surface and gradually pore structure, rich tunability and good processability, MOFs have

* Corresponding author. Department of Chemical and Biomolecular Engineering, National University of Singapore, 117585, Singapore.
** Corresponding author.
E-mail addresses: hjlin@zjnu.cn (H. Lin), chencts@nus.edu.sg (T.-S. Chung).

https://doi.org/10.1016/j.memsci.2023.121706
Received 19 January 2023; Received in revised form 20 April 2023; Accepted 29 April 2023
Available online 5 May 2023
0376-7388/© 2023 Elsevier B.V. All rights reserved.
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Fig. 1. Diagram to prepare UiO-66-NH2/CQD and TFN membranes.

attracted fast growing research interests as nanofillers [19,29–35]. ≥99%, Sinopharm Chemical Reagent), methyl blue (MB, 98%, Bide
However, because of their partial inorganic property from the metal Pharmatech), methylene blue (MLB, Sinopharm Chemical Reagent),
centers and propensity to agglomerate at a high loading, the in­ methyl orange (MO, Shanghai Zhanyun Chemical), and direct red 23
compatibility between MOFs and the polyamide matrix tends to form (DR23, Heowns Biochem Technologies) were all used without any pre­
peripheral macrovoids around MOFs and results in a defective selective treatment. Polyvinylidene difluoride (PVDF) supports were provided by
layer [19,36]. Meanwhile, although some MOFs like UiO-66-NH2 can act Da Fu Membrane Technology Co., Ltd. (Jiangsu, China). Deionized (DI)
as photocatalysts to generate reactive oxygen species (ROS), their cat­ water (resistivity: 15 MΩ cm) was obtained from a Milli-Q integral unit
alytic activity is restrained by their insufficient active sites, rapid (Millipore).
recombination of photo-induced charges and unsatisfied visible light
absorption [37–41]. 2.2. Preparation and characterizations of UiO-66-NH2/CQD
Herein, heterojunctions prepared by functionalizing UiO-66-NH2
with ultrafine carbon quantum dots (CQDs, 2− 3 nm) via covalent UiO-66-NH2 were prepared via a solvothermal process [30,45]. ZrCl4
bonding between amine groups of UiO-66-NH2 and carboxyl groups of (0.32 g) was first dissolved in 160 mL DMF and ultrasonicated for 30
CQDs were investigated as novel nanofillers in thin-film nanocomposite min. Then 0.25 g of ATA and 1.2 mL of AA were added in the solution,
(TFN) NF membranes (Fig. 1). It is believed that the carbonaceous and followed by carrying out the reaction in an oven at 100 ◦ C for 18 h.
hydrophilic CQDs could not only facilitate the nanofiller dispersion and Afterwards, the solution was centrifuged at 8000 rpm, washed by DMF
improve their compatibility in the selective layer but also enhance the and dried at 60 ◦ C. To synthesize CQDs, citric acid powders were grinded
photocatalytic property to absorb visible light and suppress recombi­ and then put in a glass container. After being heated at 180 ◦ C for 3 h,
nation of photoinduced electrons/holes [42–44]. Experimental results the white powder turned into yellow solids [46]. Then the solids were
from this work have confirmed our hypothesis. The UiO-66-NH2 nano­ dissolved in water and purified with a dialysis membrane (2K MWCO),
particles anchored with CQDs not only endowed the TFN membranes followed by being freeze dried to get CQD powders. Finally, CQDs (0.2
with higher permselectivity against dyes but also improved the water wt%) were applied to a UiO-66-NH2 aqueous solution (0.1 wt%) and
permeance from 12.4 to 23.5 L m− 2 h− 1 bar− 1. Furthermore, superior magnetically stirred for 24 h to integrate UiO-66-NH2 with CQDs, with
photocatalytic self-cleaning ability has been demonstrated by a high the presence of EDC/NHS (0.1 and 0.3 wt%, respectively). The solution
recovery rate (92-99%) of water flux after visible light irradiation for was centrifuged at 8,000 rpm and washed by DI water for several times
only 10 min. This type of novel TFN membranes may have tremendous to remove excessive CQDs and EDC/NHS.
potentials in practical wastewater treatment and purification due to the Transmission electron microscopy (TEM) images of nanoparticles
combination of outstanding separation efficiency and excellent were examined by a JEM-2100F electron microscope (accelerating
self-cleaning functionality. voltage: 200 kV). X-ray photoelectron spectroscopy (XPS) character­
izations were analyzed by an Escalab 250Xi with a monochromatized Al
2. Materials and methods cathode (1486.71 eV photons) used as the X-ray source. Core-level
spectra were retrieved with a photoelectron take-off angle of 90◦ (α,
2.1. Materials relative to the sample surface). All binding energy peaks were calibrated
according to the reference of neutral C 1s hydrocarbon peak at 284.6 eV.
2-aminoterephthalic acid (ATA, >98%, Shanghai Aladdin Biochem­ The X-ray diffraction (XRD) patterns were obtained by an X-ray
ical Technology), acetic acid (AA, Sinopharm Chemical Reagent), 1,3,5- diffractometer (Bruker D8 ADVANCE) with Cu K-α (λ = 0.154 nm) as the
bezeneetricarbonyl trichloride (TMC, 98%, Alfa Aesar), citric acid radiation source. Their surface area and porosity were determined via a
(>99.5%, Shanghai Aladdin Biochemical Technology), N-(3-dimethy­ Nova 400E by the Brunauer-Emmett-Teller (BET) method. Functional
laminopropyl)-N′ -ethylcarbodiimide hydrochloride (EDC, 98.5%, groups were investigated utilizing a Fourier transform infrared (FT-IR)
Shanghai Macklin Biochemical), piperazine (PIP, >98%, TCI Shanghai), spectroscopy (Nexus 670) with the scanning wavelength ranging from
hexanes (≥97%, Chengdu Kelong Chemical), dimethylformamide (DMF, 400 to 4000 cm− 1.
≥99.5%, Tianjin Kermel Chemical Reagent), zirconium chloride (ZrCl4,
98%, Meryer (Shanghai) Chemical Technology), N-hydroxysuccinimide 2.3. Preparation and characterizations of membranes
(NHS, 98%, Shanghai Macklin Biochemical), sodium chloride (NaCl,
≥99.5%, Shanghai Lianshi Chemical Reagent), sodium sulfate (Na2SO4, The pristine TFC NF membranes were fabricated by interfacial

2
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Fig. 2. TEM images of (a) UiO-66-NH2 (b) CQDs and (c) UiO-66-NH2/CQD; and their (d) FT-IR, (e) XRD spectra, and (f) nitrogen adsorption/desorption isotherms of
UiO-66-NH2 and their CQD composites.

polymerization of PIP and TMC as reported previously [47]. Firstly, a Na2SO4 (1000 ppm) and four representative dyes (i.e., MLB, MO,
piece of PVDF substrate was rinsed in water (containing 0.2 wt% PIP) for DR23 and MB, 5–30 mg/L) were employed to evaluate the membrane
2 min, which was then placed on a tissue paper until the surface was selectivity. The rejection (R, %) was calculated using the equation
clear of water droplet. Afterwards, it was dip-coated in hexane con­ below:
taining 0.15 wt% TMC for 1 min. Finally, the thus formed TFC mem­
brane further cured at 60 ◦ C in oven for 5 min and stored in DI water for R = 100% × (1− Cp/Cf) (2)
characterizations and performance evaluation. where Cp and Cf are the concentrations of salts or dyes in the respective
Similarly, TFN membranes were fabricated by the above-mentioned permeate and feed solutions. The Na2SO4 concentration was determined
procedures except pre-mixing the nanoparticles (0.05, 0.1, and 0.2 wt%) according to the conductivity tested on a conductivity meter (Precision
in the PIP solution. The resultant membranes were named as TFN-M (i. Scientific Instrument, Rex DDS-307A), while for the dyes, a UV–vis
e., pre-mixing with UiO-66-NH2) or TFN-MC (i.e., pre-mixing with UiO- spectrophotometer (Yipu Instrument, TU-1810) was adopted to measure
66-NH2/CQD), and the following numbers 1, 2, and 3 indicate the their concentrations.
nanoparticle concentrations in PIP solutions of 0.05, 0.1, and 0.2 wt%,
respectively. For example, TFN-MC3 is dosed with 0.2 wt% UiO-66-
NH2/CQD. 2.5. Photocatalytic activity and self-cleaning performance evaluation
The morphologies of membranes were analyzed by a field emission
scanning electron microscope (FESEM, GeminiSEM 300). Atomic force Characterizations of photocatalytic ability of UiO-66-NH2 prior and
microscopy (AFM, Bruker Dimension Icon) analyzed the surface after the CQD integration were conducted. Their ability to harvest
microstructure and quantitatively determined their average mean visible light were investigated and compared by UV–vis diffusion
roughness (Ra) and root mean square roughness (Rq) with a scanning reflectance spectra (DRS, Agilent Cary 5000). Additionally, the electron
area of 10 × 10 μm2 at 10 different places via the tapping mode in air. spin resonance (ESR) spectra of DMPO-•OH and DMPO-•O−2 were
The water contact angles along time were recorded by a goniometer recorded on a Bruker A300 spectrometer. The self-cleaning ability of
(Biolin Scientific Attention Theta Flow). Surface charges were analyzed membranes after the addition of UiO-66-NH2/CQD was evaluated using
by a zeta potential analyzer (Zeta 90 Plus). MB as the model foulant. The flux change and rejection to MB were
monitored along the time. After each cycle of 5 h, the membrane was
exposed to the visible light (provided by a 300 W Xenon lamp equipped
2.4. Evaluation of NF performance with a UV-cut filter) for 10 min. After simple rinse by DI water, the
membrane was placed back in the filtration device and continuously
Pure water permeance (PWP, L m− 2 h− 1 bar− 1) and rejections to tested for 0.5 h to determine its pure water permeance and recovery rate.
Na2SO4 and dyes (R, %) of membranes were tested using a dead-end Ten cycles were performed for each membrane sample.
filtration device at 5 bar while the membrane area was 12.6 cm2. It
was equipped with a magnetic stirrer, a digital balance and a pressurized 3. Results and discussion
N2 cylinder. In each test, the data was collected after the membrane was
stabilized for 30 min. PWP was calculated according to equation (1) as 3.1. Characterizations of UiO-66-NH2/CQD
below:

PWP = ΔV/(Am × Δt × ΔP) (1) Fig. 2(a–c) show the TEM images of UiO-66-NH2, CQDs and UiO-66-
NH2/CQD nanoparticles. After immobilization of ultrafine CQDs
where ΔV (L) represents the permeate volume through the filtration area (average size: 2–3 nm) on the UiO-66-NH2, one can see that black dots
Am (m2) during the test duration Δt (h) under a trans-membrane pressure are uniformly dispersed on the MOF surface (Fig. 2c). The FT-IR spectra
difference of ΔP (bar). shown in Fig. 2d indicate that characteristic peaks for –NH2 of UiO-66-

3
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Fig. 3. XPS spectra of UiO-66-NH2, CQDs and UiO-66-NH2/CQD.

Fig. 4. SEM (both topology and cross-section) and AFM images of membranes.

4
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Fig. 5. (a) Contact angles, (b) zeta-potential, and (c) PWP and rejections to NaCl and Na2SO4 of membranes.

NH2 (1050 and 3452 cm− 1) almost disappear in the spectra of UiO-66- confinement effect because the MOF interlayer would prevent the
NH2/CQD as they have reacted with CQDs. Meanwhile, a shoulder peak degassed nanobubbles from escape during the interfacial polymeriza­
at 1685 cm− 1 for –C– – O bond appears due to the introduction of CQDs tion, this would lead to form a thinner but larger polyamide leaves with
which are rich in these groups. XPS results also validate the existence of such a ridge-and-valley structure [48]. When embedding
–COOH on the CQD surface and their successful covalent immobilization UiO-66-NH2/CQD, the resultant TFN membrane possesses a leaf-like
on the MOF surface (Fig. 3). The high intensity of the O–C– – O peak in polyamide. However, its surface roughness is lower (i.e., Ra: 23.4 ±
CQDs’ C 1s core-level spectrum implies rich carboxyl groups on the 5.7 nm, Rq: 32.1 ± 7.8 nm) than TFN-M3 possibly due to better affinity
surface [46]. Although part of them reacted with –NH2, the leftovers still between UiO-66-NH2/CQD and PIP that decelerates the PIP diffusion for
endow the UiO-66-NH2/CQD with a higher O–C– – O peak. In addition, interfacial polymerization [45]. Both nanofillers can be easily observed
the total intensity of N 1s peak in UiO-66-NH2/CQD is weakened by the from the cross-sections of their polyamide layers.
coverage of CQDs which have no N element. As part of –NH2 has reacted Surface wetting behavior of membranes were indicated by the water
with –COOH, and part of N–H has turned into C–N, the ratio between contact angles (Fig. 5a). The TFC, TFN-M3 and TFC-MC3 membranes
N–H and C–N peaks in the N 1s spectrum of UiO-66-NH2/CQD is lower exhibit similar and relatively low water contact angles of around 69◦ at
than UiO-66-NH2. All these data confirm that CQDs are successfully the instant when the water droplet hits the membrane surfaces. How­
anchored on UiO-66-NH2. ever, the variations of the measured contact angles with time are slower
The crystallinity of UiO-66-NH2 and their CQD nanocomposites by for TFN-M3 and TFN-MC3 than TFC because the partial surface pores of
XRD were shown in Fig. 2e. All characteristic diffraction peaks of UiO- the former have been covered by the nanofillers with a rougher surface
66-NH2 are well in line with the simulated ones. The peaks at 7.3, 8.4 than the latter. Fig. 5b displays the surface charges of the membranes.
and 25.6◦ are from the respective (111), (002) and (006) planes [30,45]. The pristine TFC membrane is negatively charged, while the TFN-M3
The perfectly fitted peaks between the pristine UiO-66-NH2 and the CQD membrane becomes less negative charged because of the incorpora­
integrated ones indicate that the CQD addition has negligible influence tion of positively charged UiO-66-NH2 in the selective layer. However,
on the crystal structure of UiO-66-NH2. Fig. 2f displays the N2 adsorp­ TFN-MC3 has a more negatively charged surface than TFN-M3 owing to
tion/desorption isotherms. The pristine UiO-66-NH2 exhibit a large the attachment of the negatively charged CQDs on UiO-66-NH2
surface area of 457.9 m2/g and a N2 adsorption capacity of 740 cm3/g nanoparticles.
STP. After the CQD addition, the surface area increases to 610.8 m2/g
while the N2 adsorption capacity drops to 550 cm3/g STP. It is specu­
lated that the increment in surface area results from the surface-grafted 3.3. Filtration performance
CQDs. However, CQDs could also possibly block partial pores of
UiO-66-NH2, thus reducing the N2 adsorption. The membrane filtration performance (i.e., PWP and R to Na2SO4) of
the TFC, TFN-M and TFN-MC membranes with different nanoparticle
loadings were compared in Fig. 5c. The pristine TFC membrane has a
3.2. Characterizations of membranes satisfactory average PWP of 12.37 L m− 2 h− 1 bar− 1 and a Na2SO4
rejection of 99.3% at 5 bar. After adding 0.05 and 0.1 wt% UiO-66-NH2,
Fig. 4 illustrates the membrane morphology by both SEM and AFM. TFN-M1 and TFN-M2 membranes possess higher PWPs of 13.45 and
The TFC membrane exhibits a relatively smooth top surface with a low 19.43 L m− 2 h− 1 bar− 1, respectively, due to the “gutter layer” effect
surface roughness (i.e., Ra: 18.7 ± 2.1 nm, Rq: 24.6 ± 3.5 nm) measured [49]. However, a drop in Na2SO4 rejection was also observed which may
by AFM. After dosing UiO-66-NH2, the membrane begins to show a arise from the agglomeration of nanoparticles. A further increase in
typical ridge-and-valley architecture which gives rise to the surface UiO-66-NH2 dosage to 0.2 wt% in PIP, the membrane permeability
roughness to Ra of 34.8 ± 3.9 nm and Rq of 44.8 ± 5.6 nm. As illustrated drops due to the adverse influence of particle agglomeration. It increases
by a previous study, the difference in roughness results from the the membrane tortuosity and thus the resistance for water molecules to

5
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Table 1
Dye rejections of the control TFC and TFN-MC3 membranes.
Dyes MLB MO DR23 MB

Molecular structure

Charge + – – +
Molecular weight 319.85 327.33 769.76 799.80
R of TFC 71.94 86.75 99.81 99.66
R of TFN-MC3 82.23 96.10 99.82 99.95

Fig. 6. Rejections and selectivity for (a) Na2SO4 and dye mixed solutions and (b) NaCl and dye mixed solutions of the TFN-MC3 membranes.

Fig. 7. (a) UV–vis DRS, and ESR spectra of (b) DMPO-•OH and (c) DMPO-•O−2 of UiO-66-NH2, CQD and UiO-66-NH2/CQD; (d) possible photodegradation mech­
anism of UiO-66-NH2/CQD under visible light.

6
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Fig. 8. The normalized water fluxes and MB rejections of (a) TFC and (b) TFN-MC3 membranes as a function of time with photodegradation under visible light and
water wash after every 5 h for 10 cycles; the optical surface photographs of (c) TFC and (d) TFN-MC3 membranes before and after photodegradation and water wash.

traverse through. The growing agglomeration also reduces the Na2SO4 the Na2SO4 would shield the electrostatic repulsion between the nega­
rejection to 92.5%. In contrast, with the aid of CQD deposition on tively charged dye molecules (i.e., MO and DR23) and the membrane
UiO-66-NH2, it prevents them from agglomeration and improves the surface. Due to the ionic interactions between Na2SO4 and these two
integrity of the selective layer. Consequently, the TFN-MC3 membrane dyes, the transportation of Na2SO4 through the membranes were also
shows an enhanced PWP up to 23.46 L m− 2 h− 1 bar− 1 and an acceptable promoted, leading to a lower Na2SO4 rejection in these two cases. On the
R to Na2SO4 of 97.7%. The rejections to NaCl of membranes followed a contrary, also due to the change in charge properties of dye molecules,
similar trend of R to Na2SO4. the rejections towards the positively charged dyes (MLB and MB) in the
Both the control TFC and TFN-MC3 membranes were also evaluated mixed solutions increased, together with a higher Na2SO4 rejection
for dye removal using four commercially available dyes. Table 1 pro­ because of their retarded transportation through the membranes [47].
vides the details of these dyes (i.e., charge, molecular weight, and The selectivity between dyes and Na2SO4 is not high since their
chemical structure). Comparing to the TFC membrane, TFN-MC3 ex­ respective rejections of TFN-MC3 are both quite high. For the dye and
hibits higher rejections to all the dyes. Obviously, the addition of NaCl mixed solution, the rejection followed the same trend, while the
molecule-sieving UiO-66-NH2/CQD endows the resultant membrane selectivity between dyes and NaCl are higher than those of Na2SO4/dye
with additional size-sieving capability. In particular, even for dyes with mixtures especially for NaCl/MB (Fig. 6b).
relatively lower molecular weights like MLB and MO, TFN-MC3 can still
possess quite high rejections of 82.23 and 96.10%, respectively. The
3.4. Evaluation of self-cleaning performance
higher MO rejection is partially because of the electrostatic repulsion as
the membrane is also negatively charged. While for dyes with higher
A UV–vis DRS analyzed the light absorption abilities of CQDs, UiO-
molecular weights (i.e., DR23 and MB), TFN-MC3 has their rejections up
66-NH2, and their nanocomposites (Fig. 7a). UiO-66-NH2 has relatively
to 99.82 and 99.66%, respectively, indicating its excellent molecular
strong absorption only in the UV range (250–420 nm), consistent with
sieving property. In short, the TFN-MC3 membrane exhibits noteworthy
previous reports [50,51]. In contrast, CQDs show much stronger light
enhancement in both membrane permeability and selectivity.
harvesting especially in the visible region (400− 700 nm). As a result,
The separation performance for mixed solutions containing both
after CQD decoration, the absorption edges of the constructed nano­
salts and dyes of the TFN-MC3 membranes were also investigated. As
composites shifted to the longer wavelength with relative to
shown in Fig. 6a, compared with sole dye solutions, with the presence of
UiO-66-NH2 (Fig. 7a), implying a stronger response to visible light and
Na2SO4, the rejections against MO and DR23 are lower. This is because
more electrons/holes formed.

7
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

Fig. 7b and c compare their ROS generation using a DMPO trapped Author statement
ESR technique. The signals of DMPO-•OH and DMPO-•O−2 in the ESR
spectra confirm the generation of both •OH and •O−2 under visible light Die Ling Zhao: Methodology, Validation, Investigation, Writing-
illumination. In addition, the UiO-66-NH2/CQD nanocomposites have Original Draft. Haiyi Jin: Validation, Investigation. Qipeng Zhao: Visu­
stronger signal intensities for both •OH and •O−2 than the UiO-66-NH2. alization, Investigation. Yanchao Xu: Writing- Reviewing and Editing.
Interestingly, •OH shows a more increment in signal intensity than •O−2 , Liguo Shen: Writing- Reviewing and Editing. Hongjun Lin: Supervision.
indicating CQD facilitates more generation of •OH. These results Tai-Shung Chung:Writing- Reviewing and Editing, Supervision.
strongly validate that the addition of CQDs can improve the utilization
efficiency of visible light to produce ROS for greater photocatalytic ef­
ficiency under visible light. Fig. 7d illustrates the possible photo­ Declaration of competing interest
degradation mechanism for dyes under visible light. When the energy of
the light is equal to or higher than the energy band gap (Eg) between the The authors declare that they have no known competing financial
valence band (VB) and the conduction band (CB), the electrons of UiO- interests or personal relationships that could have appeared to influence
66-NH2/CQD nanocomposites can be excited from the VB to the reversal the work reported in this paper.
band. Then, the electrons will combine with O2 and H2O to form •O−2
and •OH, which can degrade dyes into smaller molecules and even CO2 Data availability
and H2O [52,53]. Additionally, the active radicals may inactivate mi­
croorganisms on the membrane, improving the antibacterial ability of No data was used for the research described in the article.
TFN membranes as hypothesized previously [54].
The self-cleaning ability of UiO-66-NH2/CQD embedded membranes Acknowledgments
under visible light was studied using a MB solution. Fig. 8a and b
compares the evolution of normalized water fluxes and R to MB of the This work research was supported by Zhejiang Provincial Natural
TFC and TFN-MC3 along time with photocatalysis under visible light for Science Foundation of China under Grant No. LQ23E080008. Prof.
10 min during the step of membrane washing (i.e., cleaning). The Chung likes to thank the Yushan Scholar Program supported by the
normalized water flux is the ratio of real-time water flux to the original Ministry of Education, Taiwan.
PWP after stabilization for 30 min. For the TFC membrane, the water
flux continuously declines to 82% in the first 5 h as MB molecules References
accumulated on the membrane during the filtration process. After
photocatalysis and DI water washing (Fig. 8c), the water flux is recov­ [1] M.C. van Loosdrecht, D. Brdjanovic, Anticipating the next century of wastewater
treatment, Science 344 (6191) (2014) 1452–1453.
ered to 90% due to water rinse, but it drops faster to 60% after filtrating [2] S. Xie, S. Wu, S. Bao, Y. Wang, Y. Zheng, D. Deng, L. Huang, L. Zhang, M. Lee,
the MB solution for another 5 h. After 10 consecutive runs, the PWP is Z. Huang, Intelligent mesoporous materials for selective adsorption and
only 72% of the original PWP, and water flux for dye wastewater mechanical release of organic pollutants from water, Adv. Mater. 30 (2018),
1800683.
filtration ends with 57% of the original PWP. [3] Y. He, D.L. Zhao, T.S. Chung, Na+ functionalized carbon quantum dot incorporated
For the TFN-MC3, the water flux declines to 82% in the first cycle as thin-film nanocomposite membranes for selenium and arsenic removal, J. Membr.
well (Fig. 8b). After visible light illumination for 10 min and then water Sci. 564 (2018) 483–491.
[4] J. Yuan, W.S. Hung, H. Zhu, K. Guan, Y. Ji, Y. Mao, G. Liu, K.R. Lee, W. Jin,
washing, the water flux recovery rate of the TFN-MC3 is 99%, demon­ Fabrication of ZIF-300 membrane and its application for efficient removal of heavy
strating its excellent self-cleaning ability. The optical photographs of the metal ions from wastewater, J. Membr. Sci. 572 (2019) 20–27.
membrane surface (Fig. 8d) confirm our hypothesis. The dyes on [5] R. Dai, X. Wang, C.Y. Tang, Z. Wang, Dually charged MOF-based thin-film
nanocomposite nanofiltration membrane for enhanced removal of charged
membrane surface have turned black due to photocatalytic degradation
pharmaceutically active compounds, Environ. Sci. Technol. 54 (12) (2020)
and they can be easily washed away. The PWPs in the subsequent cycles 7619–7628.
are all restored to >92% after the same procedure. It is noteworthy that [6] E. Routoula, S.V. Patwardhan, Degradation of anthraquinone dyes from effluents: a
review focusing on enzymatic dye degradation with industrial potential, Environ.
both TFC and TFN-MC membranes are robust. They have high and stable
Sci. Technol. 54 (2020) 647–664.
MB rejections before and after 10 cycles of photodegradation and water [7] C.R. Holkar, A.J. Jadhav, D.V. Pinjari, N.M. Mahamuni, A.B. Pandit, A critical
wash processes. The rejection to Na2SO4 after 10 cycles has also been review on textile wastewater treatments: possible approaches, J. Environ. Manag.
tested, which (97.93%) showed negligible difference to the original 182 (2016) 351–366.
[8] Y.C. Xu, Z.X. Wang, X.Q. Cheng, Y.C. Xiao, L. Shao, Positively charged
value (97.78%). This implies that no damage of the membrane structure nanofiltration membranes via economically mussel-substance-simulated co-
was formed after multiple photocatalysis. deposition for textile wastewater treatment, Chem. Eng. J. 303 (2016) 555–564.
[9] J. Zheng, R. Zhao, A.A. Uliana, Y. Liu, D. de Donnea, X. Zhang, B. Van der Bruggen,
Separation of textile wastewater using a highly permeable resveratrol-based loose
4. Conclusions nanofiltration membrane with excellent anti-fouling performance, Chem. Eng. J.
434 (2022), 134705.
To conclude, novel photocatalytic self-cleaning NF membranes were [10] W.M. Yu, Y. Liu, Y.C. Xu, R.J. Li, J.R. Chen, B.Q. Liao, L.G. Shen, H.J. Lin,
A conductive PVDF-Ni membrane with superior rejection, permeance and
constructed by intercalating UiO-66-NH2/CQD nanocomposites into the antifouling ability via electric assisted in-situ aeration for dye separation,
selective layer. The combination of UiO-66-NH2 and CQDs can not only J. Membr. Sci. 581 (2019) 401–412.
remarkably improve the pure water permeance by 89.7% but also in­ [11] H.B. Park, J. Kamcev, L.M. Robeson, M. Elimelech, B.D. Freeman, Maximizing the
right stuff: the trade-off between membrane permeability and selectivity, Science
crease rejections to dyes with different molecular weights and charges.
356 (2017) 1137.
In addition, CQDs enhance the visible light induced photocatalysis ac­ [12] X. Zhang, J. Ma, J. Zheng, R. Dai, X. Wang, Z. Wang, Recent advances in nature-
tivity of UiO-66-NH2 by increasing the absorption rate of visible light inspired antifouling membranes for water purification, Chem. Eng. J. 432 (2022),
134425.
and inhibiting recombination of photo-induced electrons/holes, facili­
[13] Y. Liu, L. Shen, Z. Huang, J. Liu, Y. Xu, R. Li, H. Lin, A novel in-situ micro-aeration
tating the self-cleaning process of the membrane surface by fast photo­ functional membrane with excellent decoloration efficiency and antifouling
catalytic degradation of foulants in 10 min. The water flux can be performance, J. Membr. Sci. 641 (2022), 119925.
recovered to 92− 99% of its initial state in the 10 runs of dye wastewater [14] A. Vanangamudi, D. Saeki, L.F. Dumée, M. Duke, T. Vasiljevic, H. Matsuyama,
X. Yang, Surface-engineered biocatalytic composite membranes for reduced
filtration. In contrast, by the same procedure the PWP of TFC mem­ protein fouling and self-cleaning, ACS Appl. Mater. Interfaces 10 (32) (2018)
branes can only be recovered to 72% after 10 cycles. This work may 27477–27487.
render useful opportunities for developing high-performance and self- [15] X. Jia, H. Ji, G. Zhang, J. Xing, S. Shen, X. Zhou, S. Sun, X. Wu, D. Yu, I. Wyman,
Smart self-cleaning membrane via the blending of an upper critical solution
cleaning NF membranes to treat wastewater and reclaim clean water. temperature diblock copolymer with PVDF, ACS Appl. Mater. Interfaces 13 (32)
(2021) 38712–38721.

8
D.L. Zhao et al. Journal of Membrane Science 679 (2023) 121706

[16] L.F. Dumée, J.W. Maina, A. Merenda, R. Reis, L. He, L. Kong, Hybrid thin film [36] W. Lau, S. Gray, T. Matsuura, D. Emadzadeh, J.P. Chen, A. Ismail, A review on
nano-composite membrane reactors for simultaneous separation and degradation polyamide thin film nanocomposite (TFN) membranes: history, applications,
of pesticides, J. Membr. Sci. 528 (2017) 217–224. challenges and approaches, Water Res. 80 (2015) 306–324.
[17] D.L. Zhao, Q. Zhao, H. Lin, S.B. Chen, T.S. Chung, Pressure-assisted polydopamine [37] J.H. Liu, L.G. Shen, H.J. Lin, Z.Y. Huang, H.C. Hong, C. Chen, Preparation of Ni@
modification of thin-film composite reverse osmosis membranes for enhanced UiO-66 incorporated polyethersulfone (PES) membrane by magnetic field assisted
desalination and antifouling performance, Desalination 530 (2022), 115671. strategy to improve permeability and photocatalytic self-cleaning ability, J. Colloid
[18] Q. Zhao, D.L. Zhao, S.B. Chen, T.S. Chung, Thin-film nanocomposite reverse Interface Sci. 618 (2022) 483–495.
osmosis membranes incorporated with citrate-modified layered double hydroxides [38] G. Wang, C.T. He, R. Huang, J. Mao, D. Wang, Y. Li, Photoinduction of Cu single
(LDHs) for brackish water desalination and boron removal, Desalination 527 atoms decorated on UiO-66-NH2 for enhanced photocatalytic reduction of CO2 to
(2022), 115583. liquid fuels, J. Am. Chem. Soc. 142 (45) (2020) 19339–19345.
[19] D.L. Zhao, F. Fen, L. Shen, Z. Huang, Q. Zhao, H. Lin, T.S. Chung, Engineering [39] D. Mukherjee, B. Van der Bruggestn, B. Mandal, Advancements in visible light
metal–organic frameworks (MOFs) based thin-film nanocomposite (TFN) responsive MOF composites for photocatalytic decontamination of textile
membranes for molecular separation, Chem. Eng. J. 454 (2023), 140447. wastewater: a review, Chemosphere 295 (2022), 133835.
[20] Q. Zhao, D.L. Zhao, S.B. Chen, T.S. Chung, Nanovoid-enhanced thin-film composite [40] S. Wan, M. Ou, Q. Zhong, X. Wang, Perovskite-type CsPbBr3 quantum dots/UiO-66
reverse osmosis membranes using ZIF-67 nanoparticles as a sacrificial template, (NH2) nanojunction as efficient visible-light-driven photocatalyst for CO2
ACS Appl. Mater. Interfaces 13 (2021) 33024–33033. reduction, Chem. Eng. J. 358 (2019) 1287–1295.
[21] R. Dai, H. Guo, C.Y. Tang, M. Chen, J. Li, Z. Wang, Hydrophilic selective [41] M. Xu, X. Feng, X. Han, J. Zhu, J. Wang, B. Van der Bruggen, Y. Zhang, MOF
nanochannels created by metal organic frameworks in nanofiltration membranes laminates functionalized polyamide self-cleaning membrane for advanced loose
enhance rejection of hydrophobic endocrine-disrupting compounds, Environ. Sci. nanofiltration, Sep. Purif. Technol. 275 (2021), 119150.
Technol. 53 (23) (2019) 13776–13783. [42] S. Sharma, V. Dutta, P. Singh, P. Raizada, A. Rahmani-Sani, A. Hosseini-
[22] F. Li, T.D. Liu, S. Xie, J. Guan, S. Zhang, 2D metal-organic framework-based thin- Bandegharaei, V.K. Thakur, Carbon quantum dot supported semiconductor
film nanocomposite membranes for reverse osmosis and organic solvent photocatalysts for efficient degradation of organic pollutants in water: a review,
nanofiltration, ChemSusChem 14 (11) (2021) 2452–2460. J. Clean. Prod. 228 (2019) 755–769.
[23] H. Furukawa, K.E. Cordova, M. O’Keeffe, O.M. Yaghi, The chemistry and [43] H. Nie, K. Wei, Y. Li, Y. Liu, Y. Zhao, H. Huang, M. Shao, Y. Liu, Z. Kang, Carbon
applications of metal-organic frameworks, Science 341 (6149) (2013), 1230444. dots/Bi2WO6 composite with compensatory photo-electronic effect for overall
[24] M.S. Denny, J.C. Moreton, L. Benz, S.M. Cohen, Metal–organic frameworks for water photo-splitting at normal pressure, Chin. Chem. Lett. 32 (2021) 2283–2286.
membrane-based separations, Nat. Rev. Mater. 1 (12) (2016) 1–17. [44] Q. Mou, X. Wang, Z. Xu, P. Zul, E. Li, P. Zhao, X. Liu, H. Li, G. Cheng, A synergy
[25] J. Li, H. Wang, X. Yuan, J. Zhang, J.W. Chew, Metal-organic framework establishment by metal-organic framework and carbon quantum dots to enhance
membranes for wastewater treatment and water regeneration, Coord. Chem. Rev. electrochemical water oxidation, Chin. Chem. Lett. 33 (2022) 562–566.
404 (2020) (2020), 213116. [45] D.L. Zhao, Q. Zhao, T.S. Chung, Fabrication of defect-free thin-film nanocomposite
[26] M. Kadhom, B. Deng, Metal-organic frameworks (MOFs) in water filtration (TFN) membranes for reverse osmosis desalination, Desalination 516 (2021),
membranes for desalination and other applications, Appl. Mater. Today 11 (2018) 115230.
(2018) 219–230. [46] D.L. Zhao, S. Das, T.S. Chung, Carbon quantum dots grafted antifouling membranes
[27] Y. Cheng, Y. Ying, S. Japip, S. Jiang, T.S. Chung, S. Zhang, D. Zhao, Advanced for osmotic power generation via pressure retarded osmosis process, Environ. Sci.
porous materials in mixed matrix membranes, Adv. Mater. 30 (2018), 1802401. Technol. 51 (2017) 14016–14023.
[28] N. Rangnekar, N. Mittal, B. Elyassi, J. Caro, M. Tsapatsis, Zeolite membranes–a [47] M. Xia, W. Zhang, Y. Xu, H. Lin, Y. Jiao, L. Shen, R. Li, M. Zhang, H. Hong,
review and comparison with MOFs, Chem. Soc. Rev. 44 (20) (2015) 7128–7154. Polyamide membranes with a ZIF-8@ Tannic acid core-shell nanoparticles
[29] Zhu, L. Qin, A. Uliana, J. Hou, J. Wang, Y. Zhang, X. Li, S. Yuan, J. Li, M. Tian, interlayer to enhance nanofiltration performance, Desalination 541 (2022),
J. Lin, B. Van der Bruggen, Elevated performance of thin film nanocomposite 116042.
membranes enabled by modified hydrophilic MOFs for nanofiltration, ACS Appl. [48] Y. Wen, X. Zhang, X. Li, Z. Wang, C.Y. Tang, Metal–organic framework nanosheets
Mater. Interfaces 9 (2) (2017) 1975–1986. for thin-film composite membranes with enhanced permeability and selectivity,
[30] D.L. Zhao, W.S. Yeung, Q. Zhao, T.S. Chung, Thin-film nanocomposite membranes ACS Appl. Nano Mater. 3 (9) (2020) 9238–9248.
incorporated with UiO-66-NH2 nanoparticles for brackish water and seawater [49] Z. Yang, P.F. Sun, X. Li, B. Gan, L. Wang, X. Song, H.D. Park, C.Y. Tang, A critical
desalination, J. Membr. Sci. 604 (2020), 118039. review on thin-film nanocomposite membranes with interlayered structure:
[31] M.Y. Zhang, X.P. Wang, R. Lin, Y. Liu, F.S. Chen, L.S. Cui, X.M. Meng, J. Hou, mechanisms, recent developments, and environmental applications, Environ. Sci.
Improving the hydrostability of ZIF-8 membrane by biomolecule towards enhanced Technol. 54 (24) (2020) 15563–15583.
nanofiltration performance for dye removal, J. Membr. Sci. 618 (2021), 118630. [50] Y. Pan, X. Yuan, L. Jiang, H. Wang, H. Yu, J. Zhang, Stable self-assembly AgI/UiO-
[32] S.Y. Fang, P. Zhang, J.L. Gong, L. Tang, G.M. Zeng, B. Song, W.C. Cao, J. Li, J. Ye, 66(NH2) heterojunction as efficient visible-light responsive photocatalyst for
Construction of highly water-stable metal-organic framework UiO-66 thin-film tetracycline degradation and mechanism insight, Chem. Eng. J. 384 (2020),
composite membrane for dyes and antibiotics separation, Chem. Eng. J. 385 123310.
(2020), 123400. [51] J. Pan, L. Wang, Y. Shi, L. Li, Z. Xu, H. Sun, F. Guo, W. Shi, Construction of
[33] Q. Zhao, D.L. Zhao, T.S. Chung, Thin-film nanocomposite membranes incorporated nanodiamonds/UiO-66-NH2 heterojunction for boosted visible-light photocatalytic
with defective ZIF-8 nanoparticles for brackish water and seawater desalination, degradation of antibiotics, Sep. Purif. Technol. 284 (2022), 120270.
J. Membr. Sci. 625 (2021), 119158. [52] S. Safa, M. Khajeh, A.R. Oveisi, R. Azimirad, Graphene quantum dots incorporated
[34] Y. Xu, Y. Xiao, W. Zhang, H. Lin, L. Shen, R. Li, Y. Jiao, B.-Q. Liao, Plant polyphenol UiO-66-NH2 as a promising photocatalyst for degradation of long-chain oleic acid,
intermediated metal-organic framework (MOF) membranes for efficient Chem. Phys. Lett. 762 (2021), 138129.
desalination, J. Membr. Sci. 618 (2021), 118726. [53] X. Chen, H. Gao, M. Yang, L. Xing, W. Dong, A. Li, H. Zheng, G. Wang, Smart
[35] Y. Xiao, W. Zhang, Y. Jiao, Y. Xu, H. Lin, Metal-phenolic network as precursor for integration of carbon quantum dots in metal-organic frameworks for fluorescence-
fabrication of metal-organic framework (MOF) nanofiltration membrane for functionalized phase change materials, Energy Storage Mater. 18 (2019) 349–355.
efficient desalination, J. Membr. Sci. 624 (2021), 119101. [54] D.L. Zhao, T.S. Chung, Applications of carbon quantum dots (CQDs) in membrane
technologies: a review, Water Res. 147 (2018) 43–49.

You might also like