You are on page 1of 52

Journal Pre-proofs

Removal of heavy metal ions using a new high performance nanofiltration


membrane modified with curcumin boehmite nanoparticles

Golshan Moradi, Sirus Zinadini, Laleh Rajabi, Ali Ashraf Derakhshan

PII: S1385-8947(20)30537-4
DOI: https://doi.org/10.1016/j.cej.2020.124546
Reference: CEJ 124546

To appear in: Chemical Engineering Journal

Received Date: 15 October 2019


Revised Date: 31 January 2020
Accepted Date: 21 February 2020

Please cite this article as: G. Moradi, S. Zinadini, L. Rajabi, A. Ashraf Derakhshan, Removal of heavy metal ions
using a new high performance nanofiltration membrane modified with curcumin boehmite nanoparticles, Chemical
Engineering Journal (2020), doi: https://doi.org/10.1016/j.cej.2020.124546

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will
undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing
this version to give early visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


Removal of heavy metal ions using a new high performance nanofiltration membrane
modified with curcumin boehmite nanoparticles
Golshan Moradi1, Sirus Zinadini2,* Laleh Rajabi1, Ali Ashraf Derakhshan1

1- Polymer Research Center, Department of Chemical Engineering, College of Engineering, Razi


University, Kermanshah, Iran.

2- Environmental Research Center, Department of Applied Chemistry, Razi University,


Kermanshah, Iran.

* Corresponding author:
Tel: (+98) 8334274554
Fax: (+98) 8334274554,
Email: sirus.zeinaddini@gmail.com

1
Abstract

The PES/B-Cur membranes were prepared by incorporating boehmite nanoparticles

functionalized with curcumin (B-Cur) into PES membrane via phase inversion method. The

PES/B-Cur membranes were characterized in terms of average pore size, pore size distribution,

porosity, specific surface area, water contact angle, and zeta potential. Characterization

techniques like ATR-IR, XRD, and XPS were used to confirm the presence of B-Cur

nanoparticles in the PES/B-Cur membranes. The morphological properties of B-Cur

nanoparticles and the PES/B-Cur membranes were studied using FESEM. performance of the

membranes was also evaluated for pure water flux, antifouling behavior, and removal of heavy

metal ions (Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+). The PES/B-Cur membrane showed relatively

high pure water flux (120-140 kg/m2h) than PES membrane, due to an increase in pore size,

porosity, and hydrophilicity. Moreover, the heavy metal ion removal capability of the resulting

membranes significantly increased due to the formation of chelated metal ions at the surface of

the PES/B-Cur membrane in the presence of B-Cur nanoparticles. The Fe2+, Cu2+, pb2+, Mn2+,

Zn2+, and Ni2+ rejection measured 99.88, 98.72, 99.61, 99.31, 99.11, and 99.51% for PES

membrane containing 0.5 wt.% B-Cur nanoparticles whereas they were 15.13, 14.21, 16.43,

14.38, 15.11, and 14.98% for PES. The PES membrane containing 0.5 wt.% B-Cur

nanoparticles showed the maximum adsorption capacity of 35.01 mg/g (for Pb2+), 32.20 mg/g

for (for Ni2+), 31.12 mg/g (for Cu2+), 29.08 mg/g (for Fe2+) , 27.08 mg/g (forZn2+), and 25.32

mg/g (for Mn2+). In addition, the PES/B-Cur0.5 showed higher permeate flux and FRR than other

samples through the entire process of filtration. The reusability results for the PES/B-Cur0.5

exhibited a slight reduction in rejection of Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ by only 14.5,

15.3, 13.5, 14.4, 15.1, and 14.0%.

2
Keyword: membrane, Nanofiltration, Heavy metal, Boehmite, curcumin, Antifouling

1. Introduction

Water scarcity has become a severe problem requiring urgent solution [1]. Hereupon, many

attempts have been devoted to cautiously discharge and reuse of the treated wastewater. Heavy

metals are of serious contaminations in industrial wastewater [2]. Nickel (Ni2+), lead (Pb2+), zinc

(Zn2+), iron (Fe2+), copper (Cu2+), and manganese (Mn2+) are among the most hazardous heavy

metal ions in wastewaters which are highly toxic even at low content [3]. Heavy metal

contaminations are of global concerns due to their high toxicity even at extremely low

concentrations, non-biodegradability, and high propensity to agglomeration [4, 5]. They can also

do harmful in human body, causing diseases (e. g. bone and water Minamata diseases) [6].

Hence, increasingly careful standards have been set to control heavy metal concentration in

industrial waste, which opens up high chance for water treatment methods [7]. Numerous

approaches such as reverse osmosis (RO) [8], adsorption [9], ion exchange [10], electrodialysis

[11], bioreduction [12], electrochemical reduction [13], nanofiltration (NF) [14], and evaporation

[15] have been applied to remove heavy metal ions from wastewater. However, there are still

some limitations with most of these approaches and satisfactory regularity has not reached [16].

To overcome these defects, NF membranes have currently attracted remarkable interest owing to

its versatility, high permeation flux, and removal ratio as well as low energy demand [17].

Among the nanofiltration membranes, Polyethersulfone (PES) displays extraordinary properties

(e.g. thermal stability, chemical stability over all pH rage, mechanical stability, high rigidity) in

the separation of heavy metal ions [18]. However, the main problem associated with PES

membranes (membrane fouling caused by rejected materials) resulted in a significant decrease in

separation performance, enhancement in energy consumption, increase in membrane cleaning

3
and replacement need, additional operating and preservation cost [19, 20]. Moreover, even at

high heavy metal rejections the practical application of NF membranes has been hindered due to

low permeation flux provoked by membrane fouling [21]. Weakening the interactions between

membrane and foulants is among the promising procedure for controlling membrane fouling

[22]. Therefore, several techniques like coating, surface modification, blending with fillers or

other polymers, and plasma treatment were used to alleviate the fouling of the PES membrane

through influencing surface hydrophilicity, roughness, and charge. As a result, the development

of the PES NF membranes blended with additive for the removal of heavy metals from

wastewater has attracted much attention. For instance, Ghaemi et al. [23] prepared the PES NF

membranes blended with chitosan nanopolymer, which exhibited the permeation flux of 22

kg/m2h (1.7 times higher than neat PES) and the nitrate ion removal of 93% at blending ratio of

0.2 wt.%. Adsorption sites on the membrane structure play important role in determining heavy

metal removal, and increasing the population of adsorption sites proves to improve the

adsorption of heavy metal ions and suppress the restrictions of polymer membranes [24]. It

seems that blending is also a useful method to introduce adsorption sites or functionalities on the

membrane surface. Literature reported the use of PES-based membranes for this purpose [25-27].

For example, Mojahedi et al. [28] expressed that blending of PES membrane with adsorbent

additives like polyaniline-modified TiO2 generated adsorption sites on the membrane surface and

gave raise to heavy metal removal. In another attempt to introduce adsorption sites or

functionalities on the PES membrane surface, the metformin-GO-Fe3O4 hybrid embedded PES

membrane exhibited 92% rejections of Cu2+ which was 6.5 times higher than that of unmodified

PES membrane [25]. PES membrane incorporated with CoFeO4/CuO nanoparticles showed 98,

92, and 88% rejections of Cu2+, Ni2+, and pb2+, respectively [26]. The PES membrane coated

4
with Cu-BTC/CS showed 86% and 99% rejections of Mn2+ and Fe2+, respectively [27]. Boehmite

is among alumina compound which has been used as chemical modifier, coating layer,

absorbent, reinforcement filler in ceramic composites, catalyst, optical substance, wastewater

treatment and so on [29]. Additionally, Boehmite nanoparticles possess a cubic orthorhombic

construction comprising two sheets of octahedral chains with aluminum ions core [30]. These

nanoparticles covered with abundant –OH groups at their hydrated surface preparing hydrophilic

and reactive sites which can result in an improved separation performance [31]. As expressed by

Rinaldi et al. [32] these inorganic nanoparticles possess the greatest hydrated surface and

hydrophilicity in the group of alumina compounds. The addition of boehmite nanoparticles into

membrane matrix resulted in improving antifouling resistance and hydrophilicity owing to the

superior properties of these nanoparticles like presence of extra hydroxyl groups on their surface,

hydrophilicity and other surface properties [33]. However, boehmite nanoparticles embedded

polymer membrane showed weak efficiency for metal ion removal from wastewater [31, 34].

Hence, further studies are needed to enhance the capability of boehmite to be used as an efficient

nanofiller toward fabricating membrane for heavy metal ion removal. Curcumin is a naturally

occurring polyphenolic compound made by extracting the rhizomes of turmeric [35]. According

to literature, curcumin showed considerable antioxidant and antibacterial performance [36]. To

date, the application of curcumin for fabricating membrane with high heavy metal ion removal

has not been explored. Herein, we prepare the new type of nanofiltration membranes via

incorporating boehmite nanoparticles functionalized with curcumin (B-Cur) into PES membrane

for the removal of Fe2+, Cu2+, pb2+, Mn2+, Ni2+ and Zn2+. The B-Cur nanoparticles is a favorable

candidate of scavenger for heavy metal ions owing to the powerful interactions between oxygen

atom of ketone, ether, and phenolic hydroxyl groups on the B-Cur nanoparticles with heavy

5
metal ions through complexation. Moreover, the B-Cur nanoparticles may have good

compatibility with PES polymer due to intermolecular attractions including hydrogen bonding

between phenolic hydroxyl and carbonyl functionalities of B-Cur nanoparticles with S=O of

PES. To our knowledge, it is the first study to adopt B-Cur nanoparticle as nanofiller to fabricate

mix matrix membranes which can harvest high heavy metal ion removal and antifouling

behavior at the same time. The synthesized B-Cur nanoparticles well characterized. The surface

and cross-section morphology of the membranes were characterized using FESEM. The presence

of B-Cur nanoparticles in the membrane matrix was confirmed by FTIR, XRD, and XPS

analysis. Fe2+, Cu2+, pb2+, Mn2+, Ni2+ and Zn2+ were selected as typical heavy metal

contaminations to study the removal performance of the B-Cur/PES membrane through filtration

and batch adsorption experiments. The surface hydrophilicity, porosity, average pore size, pore

size distribution, specific surface area, surface charge, pure water permeation flux, and the

antifouling behavior of the B-Cur/PES nanofiltration were also investigated.

2. Experimental

2. 1. Materials

Polyvinylpyrrolidone (PVP) with a molecular weight of 25,000 g/mol, aluminum nitrate

[Al(NO3)3_9H2O], sodium hydroxide (NaOH), Iron(II) sulfate (FeSO4), lead(II) nitrate

(Pb(NO3)2), Copper(II) nitrate (Cu(NO3)2), Potassium permanganate (KMnO4), Nickel, and

ethanol (C2H5OH) were purchased from Merck. Zinc nitrate (Zn(NO3)2) and 3–

(chloropropyl)triethoxysilane (CPTES) were purchased from Sigma-Aldrich. Polyethersulfone

(Ultrason E 6020P with Mw=58,000g/mol) and dimethylacetamide (DMAc) as a solvent were

obtained from BASF (Germany). Polyethylene glycol (PEGs) solutes (MW ~ 600, 1000, 1500

and 2000 g/mol) were obtained from Sigma Aldrich. Turmeric powder was purchased from the

6
local grocery store in Kermanshah, Iran. All chemicals used directly without further purification.

Throughout the filtration experiments, deionized water was used.

2.2. Synthesis of B-Cur nanoparticles

The extraction of curcumin from fresh turmeric powder was carried out by ethanol as a

solvent. 40 g of fine powder of turmeric was added in 200 ml ethanol under stirring for 1 h.

Afterward, Curcumin was collected in the ethanol media by filtration. Next, boehmite

nanoparticles were prepared by a method published previously [30]. In order to functionalize the

boehmite nanoparticles with curcumin, 15 g boehmite was added to the 250 ml ethanol/water

solution (1:4 v/v) and subjected to ultrasound for 15 minutes. Then 15 ml CPTES was added to

the container under nitrogen atmosphere and later refluxed at 50°C for 9 h. Afterward, 50 ml

curcumin solution was added to the container and refluxed in 70 °C for another 14 h to obtain B-

Cur nanoparticles. The orange-colored B-Cur was centrifuged and washed with ethanol. After

drying at 70 °C for 4 h, the B-Cur nanoparticles were obtained. The chemical structure of the

synthesized B-Cur nanoparticles is depicted in Fig. 1.

Fig. 1. Chemical structure of the synthesized B-Cur nanoparticles.

7
2.3. Preparation of the B-Cur/PES membrane

The fabrication of B-Cur/PES membranes was carried out via phase inversion. Typically,

different loading of nanoparticles i.e. 0.1, 0.5, and 1 wt.% were dispersed in DMAc and then

sonicated (power=140 W and frequency 35KHz) via ultrasonocator the SONREX, DT52 H

(Bandelin Co., Germany). Then, 4.5 gr PES and 0.25 PVA, pore forming agent, were added to

the suspension of nanoparticles and stirred for 1 day. Each casting solution was degassed through

keeping in a desiccator for 4 h. Afterward, the thin layer of homogenous casting solution (200

μm) was casted on a glass plate using a manual casting knife. The glass plate attached with

polymer solution layer was immediately placed into an aqueous bath at room temperature for

approximately 1.5 min and dried for perfect solidification. The prepared membranes were

marked as PES, PES/B-Cur0.1 (B-Cur nanoparticles loading ratio of 0.1 wt.%), PES/B-Cur0.5 (B-

Cur nanoparticles loading ratio of 0.5 wt.%), and PES/B-Cur1 (B-Cur nanoparticles loading ratio

of 1 wt.%).

2.4. Characterization

The chemical feature of the synthesized B-Cur nanoparticles was evaluated by Fourier transform

infrared (FTIR) spectra using a Bruker FTIR spectrophotometer at ambient temperature. The

morphological properties of the B-Cur nanoparticles and membrane samples were observed by

Field emission scanning electron microscopy (FESEM, CamScan MV2300) with already

sputtered with gold. The size of the B-Cur nanoparticles was quantified from FESEM images

using ImageJ software. The XRD patterns were obtained with an XRD

diffractometer (Philips PW 1730, Holland) using Cu Kα (1.54056 Å) radiation. The specific

surface area of the fabricated membranes was calculated using Brunauer–Emmett–Teller (BET,

8
BELSORP-MINI II, BEL Co.) at 77 K. ATR-IR spectra of all membrane samples were recorded

with Attenuated Total Reflection-Infrared (ATR-IR) spectroscopy (Thermo-Avatar, USA). X-ray

photoelectron spectroscopic (XPS, JEOL

JPS-9000SX) study of the fabricated membranes was conducted with MgKα (1253.6 eV).

Measurement of zeta potential was carried out using Electro kinetic Analyser (EKE, Anton

Paar, Austria). KCl concentration was maintained at 0.01 M; solution pH was 5.0 and 6.0. The

contact angle of 1 μl deionized water (a probe liquid) droplets that was reported the mean value

at different locations on the membrane surface as the water contact angle of membranes. The

overall porosity (ε) of the samples was calculated by the following formula:

ε(%) = (ww ― wd)/(ρ.V) × 100 (1)

Where, Ww, Wd, ρ, and V refer to the weight of wet and dry samples in kg, water density in

kg/m3, and membrane volume in m3 (calculated from available area and thickness of the

samples). The thickness of each sample was measured by using a digital micrometer (Mitotoyo,

Japan).

It is worth noting that all overall porosity measurements were carried out 4 times and the average

values were reported.

2.5. Heavy metal ions removal tests

The filtration performance of the membranes was studied by lab-scale dead-end filtration set

up as we described elsewhere [37]. Deionized water was first circulated at the set-up. Before

taking any test, each membrane sample was pressurized at 6 bar for half an hour. Then the

operating pressure was dropped to 4 bar for filtration tests and pure water flux (J0, kg/m2h) was

measured every 5 min during 1 h filtration using the following formula:

9
M
J0 = A.t (2)

Where M is the weight of collected deionized water at the permeate side (kg), A is the available

area of membrane (m2), and t is the filtration duration (h).

Next, in order to evaluate the heavy metal removal of the membranes, a low concentration of

Fe2+, Cu2+, pb2+, Mn2+, Ni2+ and Zn2+ ions (20 mg/l) worked as feeding solution. Hence, 20 mg/l

aqueous solution from FeSO4, Cu(NO3)2, Pb(NO3)2, KMnO4, Ni, and Zn(NO3)2 were prepared

separately. For Fe2+, Ni2+, Cu2+, and pb2+ pH was adjusted to 5 and for Mn2+and Zn2+ pH was

adjusted to 6. This is because of the precipitation and low stability of FeSO4, Cu(NO3)2,

Pb(NO3)2, KMnO4, Nickel, and Zn(NO3)2 aqueous solutions at pH greater than the given value

(5 and 6). All metal ions removal tests were fulfilled at room temperature. The concentration of

each solute (Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+) in solutions were determined by atomic

absorption spectroscopy (AAS). The following equation was employed to determine the Fe2+,

Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ rejection:

R(%)= 1 ― ( Cp
)
Cf × 100 (3)

Where, Cp and Cf are the concentration of each heavy metal ion in permeate and feeding

solutions, respectively.

In order to evaluate the reusability of the PES/B-Cur membranes, the heavy metal ions

removal of the membrane with the best heavy metal removal performance (at optimum loading

ratio of B-Cur nanoparticles) was studied at 10 cycles of heavy metal ion filtration test. After

each cycle, the membrane sample was placed in 1*10-2 M EDTA solution under stirring for 1 h.

10
the membrane was then placed in deionized water and the subsequent heavy metal removal

filtration test was carried out. Each heavy metal removal filtration cycle took 1 h.

2.6. Batch adsorption tests

The batch adsorption tests were carried out to determine the adsorption capacity of the

prepared membranes for removal of Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ from aqueous

solutions. The batch adsorption tests were fulfilled in a 100 ml capacity flask containing 50 ml of

the solution of each studied heavy metal ions with concentration of 20 mg/l at 25°C. For Fe2+,

Ni2+, Cu2+, and pb2+ pH was adjusted to 5 and for Mn2+and Zn2+ pH was adjusted to 6. The pH

of each testing solution was adjusted by 1M HCL/NaOH solution. The flasks were then placed

on a shaker at 180 rpm for different time intervals from 1 to 24 h. The adsorption amount of

heavy metal ions on the membrane (mg/g) was calculated using the following equation:

(C0 ― Ce) × V
qe= M
(4)

where qe denotes the adsorption amount of heavy metal ions on the membrane (mg/g), C0 and Ce

represent the initial and equilibrium concentration of heavy metal ions (mg/l), respectively. V is

the volume of the solution (l) and M is the weight of the membrane.

2.7. Antifouling behavior and pore size distribution

To determine the flux recovery ratio of the membranes, after filtration of deionized water, the

800 ppm powder milk aqueous solution was circulated in dead-end filtration setup. After

filtration of a fouling agent (powder milk aqueous solution), the membranes were placed in

deionized water for 10 min to re-determine the pure water flux (J, kg/m2h). The flux recovery

ratio was determined using the following formula [26]:

11
FRR (%) = ( ) × 100
J
J0
(5)

In order to investigate fouling behavior of the resulting membranes in more detail, irreversible

fouling ratio (Rir), reversible fouling ratio (Rr), and total fouling ratio (Rt) were determined

through the following equations [38]:

Rir = ( ) × 100
J0 ― J
J0
(6)

Rr = ( ) × 100
J ― Jp
J0
(7)

(
Rt = 1 ― J0
Jp
) (8)

Where, J0, J, and Jp are the flux of pure water, water flux of the water-cleaned membrane, and the

flux of powder milk solution in Kg/m2h.

The effective pore size and the pore size distribution of the fabricated membranes were

studied using polyethylene glycol (PEG) rejection tests with a concentration of 200 ppm

according to the previous method [39]. Typically, the rejection tests were carried out at different

PEG molecular weights (MW) of 600, 1000, 1500, and 2000 g/mol under the operating pressure

of 4 bar at 25° C and the Stokes radius (rs, m) of the PEG solutes were calculated from their

molecular weight using the following equation:

rs = 16.37 × 10 ―12 MW0.557 (9)

Subsequently, the PEG solute rejections were plotted against Stokes radius on the log-normal

probability plots. The straight line obtained from the log-normal plotting of PEG solute rejection

against Stokes radius was used to determine the mean pore radius (μp) and standard deviation

(σp) of the membranes. To further determine the pore size distribution of the membrane with

pore radius of (rp, nm), the following equation of the probability density function was used:

12
dR(rp) 1 (lnrp ― lnμp)2
drp
=rp lnσp√2π exp( ― ) (10)
2(lnσp)2

3. Results and discussion

3.1. Characterization of B-Cur nanoparticles

Fig. 2 (a) shows the FESEM of the synthesized B-Cur nanoparticles with the size range of ~

28-220 nm. XRD and FTIR analysis were also carried out to confirm the modification of

boehmite with curcumin (Fig. 2 (b, c)). From Fig. 2 (b), the diffraction peaks at 2θ value of 18.3,

25.4, and 26.5° are the characteristic peaks of curcumin [40]. The other characteristic peaks

(Marked in yellow) are assigned to the boehmite in the B-Cur nanoparticles structure. As FTIR

results show Fig. 2 (c), the frequency peak at 835 cm-1 is related to the vibrations of the C–H

bonds of alkene groups [41]. The frequencies of 1650 and 2950 cm-1 are related to the stretching

of carbonyl and –OCH3 groups, respectively [35]. On the other hand, the peak of 1650 can be

related to the C=C alkene group in curcumin. The vibrations of –CO (aromatic ether) are

indicated by peaks at 1134 and 1277 cm-1 [42]. The peak at 3450 cm-1 is ascribed to the

stretching vibrations of phenolic hydroxyl groups related to the curcumin structure [33].

13
Fig. 2. (a) FESEM images, (b) XRD pattern, and (c) FTIR spectrum of B-Cur nanoparticles.

3.2. Characterization of the PES/B-Cur membranes

The surface morphology of the PES/B-Cur0.1, PES/B-Cur0.5, and PES/B-Cur1 was investigated

by FESEM, as illustrated in Fig. 3. The surface of PES/B-Cur membranes was relatively regular

14
with globe-shaped morphology. The presence of B-Cur nanoparticles in the PES/B-Cur

membrane matrix was proved by the existence of small particles located on the membrane

surface. As the B-Cur nanoparticles loading ratio in the casting solution increased, the number of

nanoparticles on the membrane surface increased. Particularly, some large clusters were

observed on the surface of the PES/B-Cur1 sample. The higher viscosity of casting solution at

high nanoparticles loading ratio is postulated to be the crucial reason for the denser surface of

PES/B-Cur1.

The morphological information with the PES/B-Cur and PES samples is tabulated in Table 1.

The broad pore size distributions of the PES/B-Cur and PES samples specified through PEG

solute rejection are illustrated in Fig. 4. The average effective pore size was 0.69 nm for PES,

while it increased to 0.73 and 0.82 nm for PES/B-Cur0.1 and PES/B-Cur0.5, respectively. During

the phase inversion process, more immediate demixing in the presence of the B-Cur

nanoparticles resulted in producing pores with bigger size in the membrane structure. The overall

porosity of these membranes (PES/B-Cur0.1 and PES/B-Cur0.5) was also increased. Higher

porosity could be ascribed to the higher rate of diffusion between solvent (DMAc) and coagulant

(water) during phase inversion caused by the presence of B-Cur nanoparticles, as already noted.

In contrast, at higher B-Cur nanoparticles loading ratio, the average effective pore size and

overall porosity decreased. To further interpret the effect of the B-Cur nanoparticles on porous

properties of the resulting membranes the specific surface area of the membranes was calculated

using BET method (Table 1). As anticipated, the higher specific surface area values were

obtained for the PES/B-Cur membranes.

The cross-section morphology of the PES/B-Cur and PES samples was also compared via

FESEM analysis (Fig. 5). All samples showed a sandwich structure made up of a sponge-like

15
region at the top and finger-like vertical cavities with relatively large size at the inner edge.

Along with the introduction of B-Cur nanoparticles into the membrane matrix, the size of

cavities increased for both PES/B-Cur0.1 and PES/B-Cur0.5 as consistently observed by the pore

size determination results. This is because the B-Cur nanoparticles owing to the low surface

energy (due to the relative hydrophobic functionalities available on the surface of nanoparticles)

and high surface area (see Table 1) increased the rate of non-solvent and solvent exchange

producing larger cavities [43, 44]. Moreover, B-Cur nanoparticles agglomerates located on the

macro-cavities are clearly observed in a magnified image of PES/B-Cur1 (find it in yellow).

16
Fig. 3. The surface morphology of the PES/B-Cur0.1, PES/B-Cur0.5, and PES/B-Cur1 samples.

17
Table 1 The morphological information with the PES/B-Cur and PES samples.

Membrane Porosity, ε (%) Average Standard Specific

effective pore deviation, σp surface area

size, rp (nm) (m2/g)

PES 70.1±2.0 0.69 1.40 20.1

PES/B-Cur0.1 74.2±2.4 0.73 1.43 33.2

PES/B-Cur0.5 80.2±2.9 0.82 1.35 35.4

PES/B-Cur1 75.1±2.3 0.80 1.41 31.3

18
100

PES, R2=0.9969
PES/B-Cur0.1, R2=1
Solute rejection (%)

PES/B-Cur0.5,R2=1
PES/B-Cur1, R2=1

10
0.1 1 10
Pore size (nm)

1.8
Probability density function (nm-1)

PES
1.6
PES/B-Cur0.1
1.4
PES/B-Cur0.5
1.2 PES/B-Cur1

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Pore seize (nm)

Fig. 4. (a) log-normal probability plots of solute rejections against the Stokes radius and (b) pore

size distribution of the PES/B-Cur and PES samples.

19
Fig. 5. The cross-section morphology of the PES/B-Cur and PES samples.

In order to examine the dissolution of the B-Cur nanoparticles in aqueous media from the

PES/B-Cur surface, the PES/B-Cur0.5 was subjected to 20 repeated filtration cycles of heavy

metal aqueous solution. The results obtained showed that the amount of B-Cur nanoparticles

dissolved in the media was near 1-2%. The surface FESEM images of PES/B-Cur0.5 before and

after 20 repeated filtration tests, as depicted in Fig. 6, also proved the good stability of B-Cur

20
nanoparticles in PES matrix. On the other hand, the structural stability of PES/B-Cur0.5

maintained during 20 cycles of filtration and B-Cur nanoparticles did not leach into media.

Fig. 6. The surface FESEM images of PES/B-Cur0.5 a) before and b) after 20 repeated filtration

tests.

In order to approve the existence of B-Cur nanoparticles on the PES membrane, the ATR-IR

and XRD analysis were investigated. Fig. 7 (a) shows the results of ATR-IR spectra for the PES

and PES/B-Cur membranes. The PES sample shows the absorption bands at about 1149 and

1321 cm-1, which are assigned to the symmetric and asymmetric stretching of O=S=O,

respectively [45]. The presence of a peak at wavenumber at 1105 cm-1 is related to asymmetric

stretching of C–O in the structure of PES [46]. Moreover, the peaks at near 1484 and 1577 cm-1

are related to the stretching of the C6H6 ring. It was reported that the peaks around 1478 and

1574 cm-1 are characteristic ATR-IR peaks of PES [45]. The PES/B-Cur0.1 and PES/B-Cur0.5

exhibited similar ATR-IR spectra compared with PES. In these samples due to the low

concentration of the B-Cur nanoparticles, the ATR-IR spectra did not show the characteristic

peaks of the B-Cur nanoparticles. The PES/B-Cur1 sample shows the additional peaks at 1650,

21
2950, and 3450 cm-1 correspond to the stretching of carbonyl, –OCH3, and phenolic hydroxyl

groups of the B-Cur nanoparticles, respectively indicating the existence of these nanoparticles on

the surface of PES membrane. The XRD pattern of the PES/B-Cur0.5 was compared with that of

PES sample in Fig 7 (b). The characteristic peaks of PES and B-Cur nanoparticles are observed

in the XRD pattern of PES/B-Cur0.5 sample. However, less intense XRD characteristic peaks of

the B-Cur nanoparticles are observed in XRD pattern of the PES/B-Cur0.5 sample.

Fig. 7. (a) ATR-IR spectra and (b) XRD patterns of samples.

To further analyze the existence of the B-Cur nanoparticles on membrane surface, X-ray

photoelectron spectroscopy (XPS) was used (Fig. 8 and Table 2). As shown in Fig. 8 (a), the

XPS of PES sample has four peaks that can be attributed to the O 1s, N 1s, C 1s, and S 2p [47].

22
Along with introduction of 0.5 wt.% B-Cur nanoparticles into the PES matrix, new peaks arise

for Si 2p and Al 2p at approximately 101 and 74 eV. Si peak was related to the C–Si–O which

bonded the boehmite to curcumin during the synthesis of the B-Cur nanoparticles. Al peak was

also related to the boehmite source of the B-Cur nanoparticles. As shown in Table 3, along with

the introduction of 0.5 wt.% B-Cur nanoparticle into PES matrix, the atomic percentage of O

increased from 15.23 to 38.77%. This increase was a result of the high percentages of oxygen in

both PES and B-Cur nanoparticles in PES/B-Cur0.5 sample. This results display that the B-Cur

nanoparticles was decorated on the surface of PES/B-Cur0.5 sample. The high-resolution O 1s

spectrum of the PES sample can be fitted to two peaks: O=S at 532.2 eV and O–C at 533.9 eV

(Fig. 8 (b)) [47, 48]. The XPS O 1s spectrum of the PES/B-Cur0.5 sample shows the additional

peaks associated with O–Si, O–Al, O–C, and O–H bonds belong to the B-Cur nanoparticles

structure at 531.7, 532.4, 533.8, and 534.3 eV, respectively (Fig. 8 (c)) [49, 50]. As shown in

(Fig. 8 (d)), the XPS C 1s spectrum of the PES sample was fitted to four peaks. The peaks at

binding energy of 284.7 eV was assigned to the C–C and C–H of benzene rings. The peaks at

binding energies of 286.1, 285.3, and 284.5 eV were assigned to the C–O, C–S, and C=C

species, respectively. Compared with the XPS C 1s spectrum of the PES, the peaks at binding

energy of 287.3 eV related to the C=O bond of the B-Cur nanoparticles structure is observed in

the XPS C 1s spectrum of the PES/B-Cur0.5 sample.

According to the results mentioned above, from of ATR-IR, XRD, and XPS analyses, the

existence of B-Cur nanoparticles in PES matrix was confirmed.

23
Fig. 8. (a) XPS spectra of PES and PES/B-Cur0.5 samples, (b) O 1s spectrum of PES, (c) O 1s

spectrum of PES/B-Cur0.5, (d) C 1s spectrum of PES, and (e) C 1s spectrum of PES/B-Cur0.5.

24
Table 2 Different atomic concentrations for PES and PES/B-Cur0.5 samples.

Membrane O (%) C (%) Si (%) Al (%) S (%) N (%)

PES 15.23 81.67 - - 0.52 2.28

PES/B-Cur0.5 38.77 33.28 8.22 16.30 0.73 2.70

3.3. Water contact angle, pure water flux, and zeta potential

While several factors comprising morphological characteristics (porosity, pore size, active

layer thickness) and hydrophilicity influence the permeation flux, our next curiosity was to study

the active layer thickness and hydrophilicity of the prepared membranes. The active layer

thickness of the membranes was measured using ImageJ software from five random points on the

top layer in cross-section SEM images. The active layer thickness of the PES membrane

(0.73±0.04 μm) was higher than that of in PES/B-Cur0.1 (0.42±0.03 μm), PES/B-Cur0.5

(0.38±0.02 μm), and PES/B-Cur1 (0.49±0.03 μm) membranes. Higher hydrophilicity of casting

solution results in an increased the exchange rate between solvent and non-solvent during phase

inversion process [51], leading to a narrower skin layer of the PES/B-Cur membranes. Water

contact angle and pure water flux of the PES and PES/B-Cur membranes were shown in Fig. 9.

The water contact angle decreased somewhat along with the introduction of B-Cur nanoparticles

into the membrane matrix demonstrating the small change in hydrophilicity of the PES

membrane. Typically, the PES/B-Cur0.5 possessed the lowest water contact angle of 60.0° while

it was 66.1 with PES. This is due to the presence of phenolic hydroxyl and carbonyl

25
functionalities on the surface of B-Cur nanoparticles which can participate in hydrogen bonding

with water molecules. The unreacted hydrophilic –OH groups of boehmite in the structure of B-

Cur nanoparticles can also be responsible for hydrophilicity improvement in PES/B-Cur

membranes. All the PES/B-Cur membranes exhibited higher pure water flux compared to the

PES membrane. The pure water flux of PES was 114.7 kg.m2h and it reached to 123.21 and

140.45 kg.m2h for PES/B-Cur0.1 and PES/B-Cur0.5, respectively. The permeation of the

membrane is affected by the surface hydrophilicity as well as surface porosity and membrane

material [52]. Since the materials ware same, the grater pure water flux of the PES/B-Cur

membranes could mainly attributed to an increase in pore size, porosity, and hydrophilicity. In

other words, larger pores improve accommodating water molecules through the membrane [53].

However, the pure water flux decline at 1 wt.% B-Cur nanoparticles loading ratio was observed.

This water flux mitigation at high B-Cur nanoparticles loading ratio was ascribed to the high

viscosity of casting solution and formation of B-Cur aggregates on the membrane surface.

The surface charge of the PES and the PES/B-Cur membranes was assessed via calculating

the zeta potential value at pH of 5 and 6 with 1mM KCl electrolyte. From the data listed in Table

3, all investigated membranes possessed negative zeta potential at pH of 5 and 6. The zeta

potential of PES was -7.23 and -10.31 mV at pH of 5 and 6, respectively. Along with

introduction of 0.1 wt.% B-Cur nanoparticles to the PES, the zeta potential reached to -10.42 and

-12.56 mV, at pH of 5 and 6, respectively. As discussed in the earlier section, the B-Cur

nanoparticles distributed themselves on the top surface of the membrane during the phase

inversion process, therefore exposing phenolic hydroxyl and carbonyl functionalities. These

functional groups were subjected to de-protonation in aqueous medium making the surface

potential of the PES/B-Cur membranes more negative. As expressed by Deng et al. [54] these

26
effect is increased at a higher amount of nanoparticles making the surface zeta potential of the

resulting membrane more negative. Accordingly, the zeta potential of the PES/B-Cur membrane

was increased by increasing the amount of the B-Cur nanoparticles.

160
140.45
140
123.21 120.22
114.7
120

100
kg/m2h or (°)

80 66.1 62.3 60 63.3


60

40

20

0
PES PES/B-Cur0.1 PES/B-Cur0.5 PES/B-Cur1

Pure water flux (kg/m2h) Water contact angle (°)

Fig. 9. The pure water flux and water contact angle of the PES/B-Cur and PES samples.

Table 3 The zeta potential of the PES/B-Cur and PES samples.

Membrane Zeta potential at pH=5 (mV) Zeta potential at pH=6 (mV)

PES -7.23 -10.31

PES/B-Cur0.1 -10.42 -12.56

PES/B-Cur0.5 -14.34 -15.78

PES/B-Cur1 -15.46 -16.57

27
3.4. Filtration of aqueous solutions containing heavy metal ions

3.4.1. Heavy metal ion rejection

Fig. 10 shows the heavy metal ion rejection results of the PES/B-Cur and PES membranes

during 3 h filtration under the operating pressure of 4 bar at room temperature. The PES/B-Cur0.5

showed higher rejection for all tested metal ions (rejections of Fe2+:99.88%, Cu2+:98.72%,

pb2+:99.61%, Mn2+:99.31%, Zn2+:99.11%, and Ni2+:99.51%) than PES/B-Cur0.1 (rejections of

Fe2+:87.71, Cu2+:86.75%, pb2+:87.52%, Mn2+:87.17%, Zn2+:87.98%, and Ni2+:87.51%) and

PES/B-Cur1 (rejections of Fe2+:82.82, Cu2+:81.61%, pb2+:82.53%, Mn2+:82.37%, Zn2+:82.41%,

and Ni2+:82.50%). The metal ion rejection with PES/B-Cur0.5 was about 5.5 (for Fe2+), 6.9 (for

Cu2+), 6.0 (for pb2+), 6.9 (for Mn2+), 6.5 (for Zn2+), and 6.6 (for Ni2+) times higher than that of

the control membrane (rejections of Fe2+:15.13, Cu2+:14.21%, pb2+:16.43%, Mn2+:14.38%,

Zn2+:15.11%, and Ni2+:14.98%). The significant increment in heavy metal ions rejection values

was due to the incorporation of functional groups (e.g. ketone, ether, and phenolic hydroxyl), as

confirmed by XRD, FTIR, and XPS analysis, on the membrane surface through the introduction

of B-Cur nanoparticles into the PES matrix. On the other hand, the complex formation between

the oxygen atom of ketone, ether, and phenolic hydroxyl groups on the B-Cur nanoparticles and

heavy metal ions, as shown in Fig. 11, and subsequent formation of chelated metal ions were

responsible for the adsorption of metal ions with the PES/B-Cur membrane. Besides, as zeta

potential results suggested, along with the introduction of the B-Cur nanoparticles the zeta

potential of the PES membrane became more negative. Consequently, the electrostatic exclusion

of positively charged ions was higher in the PES/B-Cur membranes compared to the PES,

resulting in the higher heavy metal ions rejection. It should be noted that the heavy metal ion

28
rejection depends on electrostatic exclusion with the membrane surface and the pore size of the

membrane. Although the PES/B-Cur membranes had bigger pore size than PES

(see Table 1), their rejection was higher. Hence, the electrostatic exclusion was dominant for the

removal of heavy metal ions for the PES/B-Cur membranes. However, Some decrement in heavy

metal ions rejection was observed for PES/B-Cur1 containing 1 wt.% B-Cur nanoparticles. This

was attributed to the possible agglomeration of nanoparticles at high content which results in the

reduced effective surface of nanoparticles and available active adsorption sites [55, 56]. It

should be noted that, in order to prove the reproducibility of the PES/B-Cur membranes, each

heavy metal ion rejection performance was the mean of at least 5 membranes prepared at the

same condition.

120

100
Fe2+ rejection (%)

80

60
PES PES/B-Cur0.5 PES/B-Cur0.1 PES/B-Cur1

40

20

0
0 50 100 150 200 250 300
Time (min)

29
120

100
Cu2+ rejection (%)
80

60
PES PES/B-Cur0.1 PES/B-Cur0.5 PES/B-Cur1
40

20

0
0 50 100 150 200 250 300
Time (min)

120

100
pb2+ rejection (%)

80

60
PES PES/B-Cur0.1 PES/B-Cur0.5 PES/B-Cur1
40

20

0
0 50 100 150 200 250 300
Time (min)

30
120

100

Mn2+ rejection (%) 80

60 PES PES/B-Cur0.1

40
PES/B-Cur0.5 PES/B-Cur1

20

0
0 50 100 150 200 250 300
Time (min)

120

100
Zn2+ rejection (%)

80
PES PES/B-Cur0.1
60

40 PES/B-Cur0.5 PES/B-Cur1

20

0
0 50 100 150 200 250 300
Time (min)

31
120

100
Ni2+ rejection (%)
80

PES PES/B-Cur0.1
60

40 PES/B-Cur0.5 PES/B-Cur1

20

0
0 50 100 150 200 250 300
Time (min)

Fig. 10. heavy metal ion (Fe2+, Cu2+, pb2+, Mn2+, Zn2+, and Ni2+) rejection results of the PES/B-

Cur and PES samples.

Fig. 11. Complexation of metal ions with functional groups on the B-Cur nanoparticles located

on the surface of the PES/B-Cur nanoparticles.


32
The regeneration investigation of the PES/B-Cur membranes is serious in order to examine

the potential for industrial applications. Hence, the reusability of PES/B-Cur0.5 (as the optimum

membrane) was tested based on Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ rejection under the same

condition (operating pressure= 4 bar and metal ion concentration in aqueous solution=20 mg/l).

10 repeated cycles of experiment were carried out and each cycle took 1 h. As shown in Fig. 12,

After the 1 h filtration of each metal ion aqueous solution, the testing membrane was immersed

in EDTA solution and filtration experiment was performed for the next 1 h. As shown in Fig. 12,

a slight loss of metal ion rejection was observed. For instance, the rejection of Fe2+ with PES/B-

Cur0.5 was, 99.98, 97.23, 95.22, 93.52, 92,43, 90.23, 88,37, 87,81, 86.53, and 85.01% for first,

second, third, …, and tenth run in 600 min filtration, respectively. The Fe2+, Cu2+, pb2+, Mn2+,

Ni2+, and Zn2+ rejection was reduced by only 14.5, 15.3, 13.5, 14.4, 15.1, and 14.0% after the

tenth run, respectively, which proved capable for reusability of the PES/B-Cur0.5 without a

tangible loosing of its heavy metal ion removal efficiency.

105
100 Fe2+ Cu2+ pb2+ Mn2+ Ni2+ Zn2+
95
90
Rejection (%)

85
80
75
70
65
60
55
50
1 2 3 4 5 6 7 8 9 10
Cycle number

33
Fig. 12. Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ rejection of the PES/B-Cur0.5 during 10 repeated

cycles of experiment.

3.4.2. Effect of foreign ions

In order to investigate the practical application of the PES/B-Cur membranes, the metal ion

removal experiment for PES/B-Cur0.5 was carried out in the presence of foreign ions of Na+ and

Mg2+. As shown in Fig. 13, it was observed that the rejections were slightly decreased from

Fe2+:99.88%, Cu2+:98.72%, pb2+:99.61%, Mn2+:99.31%, Zn2+:99.11%, and Ni2+:99.51% to

Fe2+:97.45%, Cu2+:95.45%, pb2+:97.11%, Mn2+:96.87%, Zn2+:96.12%, and Ni2+:97.00% in

presence of Na+. However, the rejections were decreased to a greater degree in presence of Mg2+

(Fe2+:73.56%, Cu2+:70.76%, pb2+:72.23%, Mn2+:71.43%, Zn2+:71.00%, and Ni2+:71.94%). This

was due to the fact that the mobility of studied metal ions (Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+)

towards the membrane surface was reduced in presence of higher-valent metal ions such as

Mg2+. Moreover, as explained by Syed Ibrahim et al. [57] complexation of foreign metal ions

with membrane surface functional groups decreases the number of active available functional

groups for Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ to bind.

34
110

Fe2+ Cu2+ pb2+ Mn2+ Zn2+ Ni2+


100

90
Rejection (%)

80

70

60

50

40
Na+ Mg2+
Metal ions

Fig. 13. Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+ rejection in presence of foreign ions of Na+ and

Mg2+ (4 bar, 20 mg/l, pH of 5 for Fe2+, Ni2+, Cu2+, and pb2+ and 6 for Mn2+and Zn2+).

3.5. Adsorption study

In this study, among the PES/B-Cur membranes at different B-Cur nanoparticles amounts, the

well-performed membrane PES/B-Cur0.5 (higher heavy metal removal during filtration) was

favored for the study of heavy metal adsorption on the membrane surface. Hence, to study the

performance of the PES/B-Cur membranes towards the Fe2+, Cu2+, pb2+, Mn2+, Ni2+, and Zn2+,

the batch adsorption test of PES/B-Cur0.5 towards these metal ions were carried out at same

initial concentration of 20 mg/l and different time intervals up to 1 day and the related results are

shown in Fig. 14. The adsorption capacity of studied metal ions onto PES and PES/B-Cur0.5

membranes increased drastically within the initial 6 h, then rose gradually and reached

35
equilibrium in about 10 h due to the saturation adsorption of active sites. The PES exhibited poor

adsorption capacity for heavy metal ions (Fe2+, Cu2+, Pb2+, Mn2+, Ni2+, and Zn2+) owing to the

absence of main functional groups on its surface. The PES/B-Cur0.5 had higher adsorption

capacity than PES verifying that the introduction of B-Cur nanoparticles on the PES membrane

indeed ameliorated the adsorption of heavy metal ions and increased the number of active sites.

In case of the PES/B-Cur0.5, the order of adsorption capacity for the studied heavy metal ions was

Pb2+ > Ni2+ > Cu2+ > Fe2+ > Zn2+ > Mn2+. The maximum amount of Pb2+, Ni2+, Cu2+, Fe2+ , Zn2+,

and Mn2+ adsorbed onto the PES/B-Cur0.5 were 35.01, 32.20, 31.12, 29.08, 27.08, and 25.32

mg/g, respectively. The increased adsorption capacity towards pb2+ was related to the lower

hydrated radii and higher electronegativity of pb2+. At higher hydrated radii the adsorption sites

are saturated faster due to a steric hindrance, while higher electronegativity induces higher

susceptibility of heavy metal ions to form a complex with the adsorbent surface [58]. The values

of hydrated radii follow the sequence [59] : Pb2+ (4.01 Å) < Ni2+ (4.04 Å) < Cu2+ (4.19 Å) < Fe2+

(4.28 Å) < Zn2+ (4.30 Å) < Mn2+ (4.38 Å), while the values of electronegativity follow the

sequence [60] : pb2+ (2.20) > Ni2+ (1.91) > Cu2+ (1.90) > Fe2+ (1.85) > Zn2+ (1.65) > Mn2+ (1.55).

Lower hydrated radii and higher electronegativity facilitate the adsorption of heavy metal ions.

36
40

35

30
Adsorption capacity (mg/g)

25

20

15

10

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26
Time (h)

Fig. 14. Adsorption capacity of Fe2+, Cu2+, Pb2+, Mn2+, Ni2+, and Zn2+ at different time intervals

on PES and PES/B-Cur0.5 membranes (initial concentration of 20 mg/l, rotation speed of 180

rpm, temperature of 25°C , pH of 5 for Fe2+, Ni2+, Cu2+, and pb2+ and 6 for Mn2+and Zn2+).

3.6. Antifouling properties of the PES/B-Cur membranes

Antifouling behavior of the nanofiltration membrane is indispensable features in long-term

feasible applications like wastewater treatment [61, 62]. The antifouling behavior of the PES/B-

Cur membranes was examined by four-cycle filtration test of powder milk solution and

compared with PES. As the permeate flux results (Fig. 15 (a)) show, all PES/B-Cur membranes

showed higher permeate flux than PES. In each cycle, the permeate flux of both PES/B-Cur and

PES were suddenly reduced when deionized water in filtration cell was switched to milk powder

solution and then attained a stable value during 15 minutes. This flux decline was due to protein

37
aggregation and pore blocking [63]. Afterward, the permeate flux remained stable with only

small fluctuations. After 240 minutes filtration (0-60 min deionized water and 60-240 min

powder milk solution) and membrane cleaning, the pure water flux of the cleaned samples were

recovered to some degree. In the case of the PES/B-Cur membranes, the recovery of pure water

flux was more satisfactory, especially with PES/B-Cur0.5. As depicted in Fig. 15 (b), in the case

of 1st cycle of filtration, the FRR of PES was only 30.10% while it was 92.12, 98.51, and

77.65% for PES/B-Cur0.1, PES/B-Cur0.5, and PES/B-Cur1, respectively. It was also found that the

FRR of the PES/B-Cur0.5 remained above 94.6% within four-cycle filtration, while the FRR of

PES was 30.10, 26.10, 25.10, 24.11% for 1st, 2nd, 3rd, and 4 cycles in 21 h filtration,

respectively. Interestingly, the PES/B-Cur0.5 showed higher permeate flux and FRR than other

samples through the entire process of filtration, indicating its higher antifouling behavior. This

improved antifouling behavior along with the introduction of B-Cur nanoparticles may be

attributed to the chemical structure of the membrane surface. Location of B-Cur nanoparticles

introduced hydroxyl and carbonyl functionalities over the PES/B-Cur surface which can

participate in hydrogen bonding with water molecules and generate a continuous layer of water

on the PES/B-Cur surface, hindering the foulant adsorption. Fig. 15 (c) provides more precise

detail on antifouling behavior of the PES/B-Cur membranes. Three indexes including the

irreversible fouling ratio (Rir ), reversible fouling ratio (Rr ), total fouling ratio (Rt ) are shown in

Fig. 15 (c). The related results revealed that the PES/B-Cur membranes had lower Rir and Rt

compared with PES sample, while Rr was higher. In the best case, both Rir and Rt values

significantly reduced from 73.43 and 84.98% (for PES) to 1.49 and 28.1% for the PES/B-Cur0.5

which was lower than the reported value for modified polymer nanofiltration membrane in the

literature [38, 64].

38
160
PES PES/B-Cur0.1 PES/B-Cur0.5 PES/B-Cur1
140

120
Flux (kg/m2h )

100

80

60

40

20

0
0 200 400 600 800 1000 1200
Time (min)

120

100

80
FRR (%)

60

40

20

0
PES PES/B-Cur0.1 PES/B-Cur0.5 PES/B-Cur1

Cycle 1 Cycle 2 Cycle 3 Cycle 4

39
180
160

Fouling resistance (%)


140
120
100
80
60
40
20
0
PES PES/B-Cur0.1 PES/B-Cur0.5 PES/B-Cur1

Ireversible fouling ratio reversible fouling ratio Total fouling ratio

Fig. 15. (a) Time-dependent permeate flux, (b) FRR values of the PES/B-Cur and PES samples

with the quaternary cycle filtration tests using powder milk solution as the model foulant, and (c)

Rir, Rr, and Rt of the PES/B-Cur and PES samples.

4. Conclusion

The effect of B-Cur nanoparticles on the morphology, heavy metal ion rejection, batch

adsorption, and antifouling behavior of the PES nanofiltration membranes was investigated. The

B-Cur nanoparticles were synthesized and characterized by FTIR, XRD, and FESEM analysis to

confirm the decisive formation of the B-Cur nanoparticles. As FESEM, pore size distribution,

and porosity measurements suggest, the presence of B-Cur nanoparticles in membrane matrix

resulted in increased pore size and higher porosity of the PES/B-Cur membranes. The water

contact angle decreased somewhat along with the introduction of B-Cur nanoparticles into

membrane matrix demonstrating the small change in hydrophilicity of the PES membrane.

Typically, the PES/B-Cur0.5 possessed the lowest water contact angle of 60.0° while it was 66.1°

40
with PES. This is due to the presence of phenolic hydroxyl and carbonyl functionalities on the

surface of B-Cur nanoparticles which can participate in hydrogen bonding with water molecules.

All the PES/B-Cur membranes exhibited higher pure water flux compared to the PES membrane.

The pure water flux of PES was 114.7 kg.m2h and it reached to 123.21 and 140.45 kg.m2h for

PES/B-Cur0.1 and PES/B-Cur0.5, respectively. Besides, the Fe2+, Cu2+, pb2+, Mn2+, Zn2+, and Ni2+

rejection measured 99.88, 98.72, 99.61, 99.31, 99.11, and 99.51% for the PES/B-Cur0.5 whereas

they were 15.13, 14.21, 16.43, 14.38, 15.11, and 14.98% for PES. The PES/B-Cur0.5 showed the

maximum adsorption capacity of 35.01 mg/g (for Pb2+), 32.20 mg/g for (for Ni2+), 31.12 mg/g

(for Cu2+), 29.08 mg/g (for Fe2+) , 27.08 mg/g (forZn2+), and 25.32 mg/g (for Mn2+). The

reusability results for the PES/B-Cur0.5 exhibited a slight reduction in rejection of Fe2+, Cu2+,

pb2+, Mn2+, Ni2+, and Zn2+ by only 14.5, 15.3, 13.5, 14.4, 15.1, and 14.0%. The PES/B-Cur

membranes exhibited more permeation flux and outstanding antifouling behavior compared to

PES during 4 cycles of filtration in 21 h. The related results revealed that the PES/B-Cur

membranes had lower Rir and Rt compared with PES sample, while Rr was higher. Thanks to the

results obtained, this type of antifouling membrane opens a new horizon for future applications.

41
References

[1] C.J. Vörösmarty, P.B. McIntyre, M.O. Gessner, D. Dudgeon, A. Prusevich, P. Green, S.

Glidden, S.E. Bunn, C.A. Sullivan, C.R. Liermann, Global threats to human water security and

river biodiversity, Nature 467 (2010) 555.

[2] G. Moradi, F. Dabirian, P. Mohammadi, L. Rajabi, M. Babaei, N. Shiri, Electrospun fumarate

ferroxane/polyacrylonitrile nanocomposite nanofibers adsorbent for lead removal from aqueous

solution: Characterization and process optimization by response surface methodology, Chemical

Engineering Research and Design 129 (2018) 182-196.

[3] A. Bera, J.S. Trivedi, S.B. Kumar, A.K.S. Chandel, S. Haldar, S.K. Jewrajka, Anti-organic

fouling and anti-biofouling poly (piperazineamide) thin film nanocomposite membranes for low

pressure removal of heavy metal ions, Journal of hazardous materials 343 (2018) 86-97.

[4] G. Crini, Non-conventional low-cost adsorbents for dye removal: a review, Bioresource

technology 97 (2006) 1061-1085.

[5] C. Blöcher, J. Dorda, V. Mavrov, H. Chmiel, N. Lazaridis, K. Matis, Hybrid flotation—

membrane filtration process for the removal of heavy metal ions from wastewater, Water

Research 37 (2003) 4018-4026.

[6] M. Athar, S.B. Vohora, Heavy metals and environment, New Age International1995.

[7] Y. Xu, D. Zhao, Removal of copper from contaminated soil by use of poly (amidoamine)

dendrimers, Environmental science & technology 39 (2005) 2369-2375.

[8] Y. Gao, W. Li, W.C. Lay, H.G. Coster, A.G. Fane, C.Y. Tang, Characterization of forward

osmosis membranes by electrochemical impedance spectroscopy, Desalination 312 (2013) 45-

51.

42
[9] G.Z. Kyzas, G. Bomis, R.I. Kosheleva, E.K. Efthimiadou, E.P. Favvas, M. Kostoglou, A.C.

Mitropoulos, Nanobubbles effect on heavy metal ions adsorption by activated carbon, Chemical

Engineering Journal 356 (2019) 91-97.

[10] A. Bashir, L.A. Malik, S. Ahad, T. Manzoor, M.A. Bhat, G. Dar, A.H. Pandith, Removal of

heavy metal ions from aqueous system by ion-exchange and biosorption methods, Environmental

Chemistry Letters 17 (2019) 729-754.

[11] S.A. Al-Saydeh, M.H. El-Naas, S.J. Zaidi, Copper removal from industrial wastewater: A

comprehensive review, Journal of industrial and engineering chemistry 56 (2017) 35-44.

[12] D. Pradhan, L.B. Sukla, M. Sawyer, P.K. Rahman, Recent bioreduction of hexavalent

chromium in wastewater treatment: a review, Journal of Industrial and Engineering Chemistry 55

(2017) 1-20.

[13] C. Barrera-Díaz, V. Lugo-Lugo, G. Roa-Morales, R. Natividad, S. Martínez-Delgadillo,

Enhancing the electrochemical Cr (VI) reduction in aqueous solution, Journal of hazardous

materials 185 (2011) 1362-1368.

[14] Y. Qi, L. Zhu, X. Shen, A. Sotto, C. Gao, J. Shen, Polythyleneimine-modified original

positive charged nanofiltration membrane: Removal of heavy metal ions and dyes, Separation

and Purification Technology 222 (2019) 117-124.

[15] R.-S. Juang, M.-N. Chen, Removal of copper (II) chelates of EDTA and NTA from dilute

aqueous solutions by membrane filtration, Industrial & engineering chemistry research 36 (1997)

179-186.

[16] L. Liu, Y. Xu, K. Wang, K. Li, L. Xu, J. Wang, J. Wang, Fabrication of a novel conductive

ultrafiltration membrane and its application for electrochemical removal of hexavalent

chromium, Journal of Membrane Science 584 (2019) 191-201.

43
[17] B. Al-Rashdi, C. Somerfield, N. Hilal, Heavy metals removal using adsorption and

nanofiltration techniques, Separation & Purification Reviews 40 (2011) 209-259.

[18] A.P. Mojdehi, M.P. Chenar, M.N. Mahboub, M. Eftekhari, Development of

PES/polyaniline-modified TiO2 adsorptive membrane for copper removal, Colloids and Surfaces

A: Physicochemical and Engineering Aspects (2019) 123931.

[19] Q. Zhang, J. Jiang, F. Gao, G. Zhang, X. Zhan, F. Chen, Engineering high-effective

antifouling polyether sulfone membrane with P (PEG-PDMS-KH570)@ SiO2 nanocomposite

via in-situ sol-gel process, Chemical Engineering Journal 321 (2017) 412-423.

[20] R. Liu, Y. Sui, X. Wang, Metal–organic framework-based ultrafiltration membrane

separation with capacitive-type for enhanced phosphate removal, Chemical Engineering Journal

371 (2019) 903-913.

[21] Y.-L. Ji, Q.-F. An, F.-Y. Zhao, C.-J. Gao, Fabrication of chitosan/PDMCHEA blend

positively charged membranes with improved mechanical properties and high nanofiltration

performances, Desalination 357 (2015) 8-15.

[22] F. Gao, G. Zhang, Q. Zhang, X. Zhan, F. Chen, Improved antifouling properties of poly

(ether sulfone) membrane by incorporating the amphiphilic comb copolymer with mixed poly

(ethylene glycol) and poly (dimethylsiloxane) brushes, Industrial & Engineering Chemistry

Research 54 (2015) 8789-8800.

[23] N. Ghaemi, P. Daraei, F.S. Akhlaghi, Polyethersulfone nanofiltration membrane embedded

by chitosan nanoparticles: fabrication, characterization and performance in nitrate removal from

water, Carbohydrate polymers 191 (2018) 142-151.

[24] K.K. Tetala, D.F. Stamatialis, Mixed matrix membranes for efficient adsorption of copper

ions from aqueous solutions, Separation and purification technology 104 (2013) 214-220.

44
[25] G. Abdi, A. Alizadeh, S. Zinadini, G. Moradi, Removal of dye and heavy metal ion using a

novel synthetic polyethersulfone nanofiltration membrane modified by magnetic graphene

oxide/metformin hybrid, Journal of membrane science 552 (2018) 326-335.

[26] F. Zareei, S.M. Hosseini, A new type of polyethersulfone based composite nanofiltration

membrane decorated by cobalt ferrite-copper oxide nanoparticles with enhanced performance

and antifouling property, Separation and Purification Technology 226 (2019) 48-58.

[27] M. Mozafari, S.F. Seyedpour, S.K. Salestan, A. Rahimpour, A.A. Shamsabadi, M.D.

Firouzjaei, M.R. Esfahani, A. Tiraferri, H. Mohsenian, M. Sangermano, Facile Cu-BTC surface

modification of thin chitosan film coated polyethersulfone membranes with improved antifouling

properties for sustainable removal of manganese, Journal of Membrane Science (2019) 117200.

[28] A.P. Mojdehi, M.P. Chenar, M. Namvar-Mahboub, M. Eftekhari, Development of

PES/polyaniline-modified TiO2 adsorptive membrane for copper removal, Colloids and Surfaces

A: Physicochemical and Engineering Aspects 583 (2019) 123931.

[29] Y. Mathieu, B. Lebeau, V. Valtchev, Control of the morphology and particle size of

boehmite nanoparticles synthesized under hydrothermal conditions, Langmuir 23 (2007) 9435-

9442.

[30] L. Rajabi, A. Derakhshan, Room temperature synthesis of boehmite and crystallization of

nanoparticles: effect of concentration and ultrasound, Science of advanced materials 2 (2010)

163-172.

[31] X. Wang, T. Wang, J. Ma, H. Liu, P. Ning, Synthesis and characterization of a new

hydrophilic boehmite-PVB/PVDF blended membrane supported nano zero-valent iron for

removal of Cr (VI), Separation and Purification Technology 205 (2018) 74-83.

45
[32] R. Rinaldi, U. Schuchardt, On the paradox of transition metal-free alumina-catalyzed

epoxidation with aqueous hydrogen peroxide, Journal of Catalysis 236 (2005) 335-345.

[33] V. Vatanpour, S.S. Madaeni, L. Rajabi, S. Zinadini, A.A. Derakhshan, Boehmite

nanoparticles as a new nanofiller for preparation of antifouling mixed matrix membranes,

Journal of membrane science 401 (2012) 132-143.

[34] G. Hota, B.R. Kumar, W. Ng, S. Ramakrishna, Fabrication and characterization of a

boehmite nanoparticle impregnated electrospun fiber membrane for removal of metal ions,

Journal of materials science 43 (2008) 212-217.

[35] Z. Song, Y. Wu, H. Wang, H. Han, Synergistic antibacterial effects of curcumin modified

silver nanoparticles through ROS-mediated pathways, Materials Science and Engineering: C 99

(2019) 255-263.

[36] M.R. Agel, E. Baghdan, S.R. Pinnapireddy, J. Lehmann, J. Schäfer, U. Bakowsky,

Curcumin loaded nanoparticles as efficient photoactive formulations against gram-positive and

gram-negative bacteria, Colloids and Surfaces B: Biointerfaces 178 (2019) 460-468.

[37] G. Moradi, S. Zinadini, F. Dabirian, L. Rajabi, Polycitrate-para-aminobenzoate alumoxane

nanoparticles as a novel nanofiller for enhancement performance of electrospun PAN

membranes, Separation and Purification Technology 213 (2019) 224-234.

[38] F. Mohammadnezhad, M. Feyzi, S. Zinadini, A novel Ce-MOF/PES mixed matrix

membrane; synthesis, characterization and antifouling evaluation, Journal of industrial and

engineering chemistry 71 (2019) 99-111.

[39] F. Soyekwo, C. Liu, H. Wen, Y. Hu, Construction of an electroneutral zinc incorporated

polymer network nanocomposite membrane with enhanced selectivity for salt/dye separation,

Chemical Engineering Journal 380 (2020) 122560.

46
[40] J. Jia, N. Song, Y. Gai, L. Zhang, Y. Zhao, Release-controlled curcumin proliposome

produced by ultrasound-assisted supercritical antisolvent method, The Journal of Supercritical

Fluids 113 (2016) 150-157.

[41] H.-J. Kim, D.-J. Kim, S. Karthick, K. Hemalatha, C.J. Raj, S. Ok, Y. Choe, Curcumin dye

extracted from Curcuma longa L. used as sensitizers for efficient dye-sensitized solar cells,

International Journal of electrochemical science 8 (2013) 8320-8328.

[42] E. Ismail, D. Sabry, H. Mahdy, M. Khalil, Synthesis and Characterization of some Ternary

Metal Complexes of Curcumin with 1, 10-phenanthroline and their Anticancer Applications,

Journal of Scientific Research 6 (2014) 509-519.

[43] D. Hou, G. Dai, J. Wang, H. Fan, L. Zhang, Z. Luan, Preparation and characterization of

PVDF/nonwoven fabric flat-sheet composite membranes for desalination through direct contact

membrane distillation, Separation and purification technology 101 (2012) 1-10.

[44] F. Ardeshiri, S. Salehi, M. Peyravi, M. Jahanshahi, A. Amiri, A.S. Rad, PVDF membrane

assisted by modified hydrophobic ZnO nanoparticle for membrane distillation, Asia‐Pacific

Journal of Chemical Engineering 13 (2018) e2196.

[45] P.G. Ingole, W. Choi, K.H. Kim, C.H. Park, W.K. Choi, H.K. Lee, Synthesis,

characterization and surface modification of PES hollow fiber membrane support with

polydopamine and thin film composite for energy generation, Chemical Engineering Journal 243

(2014) 137-146.

[46] S. Kamari, A. Shahbazi, Biocompatible Fe3O4@ SiO2-NH2 nanocomposite as a green

nanofiller embedded in PES–nanofiltration membrane matrix for salts, heavy metal ion and dye

removal: Long–term operation and reusability tests, Chemosphere 243 (2020) 125282.

47
[47] J. Wang, R. He, X. Han, D. Jiao, J. Zhu, F. Lai, X. Liu, J. Liu, Y. Zhang, B. Van der

Bruggen, High performance loose nanofiltration membranes obtained by a catechol-based route

for efficient dye/salt separation, Chemical Engineering Journal (2019) 121982.

[48] S.X. Liu, J.-T. Kim, Characterization of surface modification of polyethersulfone

membrane, Journal of adhesion science and technology 25 (2011) 193-212.

[49] A.K. Singh, S. Yadav, K. Sharma, Z. Firdaus, P. Aditi, K. Neogi, M. Bansal, M.K. Gupta,

A. Shanker, R.K. Singh, Quantum curcumin mediated inhibition of gingipains and mixed-biofilm

of Porphyromonas gingivalis causing chronic periodontitis, RSC advances 8 (2018) 40426-

40445.

[50] J. Ma, W.-J. Lee, J.M. Bae, K.-S. Jeong, S.H. Oh, J.H. Kim, S.-H. Kim, J.-H. Seo, J.-P.

Ahn, H. Kim, Carrier mobility enhancement of tensile strained Si and SiGe nanowires via

surface defect engineering, Nano letters 15 (2015) 7204-7210.

[51] L. Bai, H. Wu, J. Ding, A. Ding, X. Zhang, N. Ren, G. Li, H. Liang, Cellulose nanocrystal-

blended polyethersulfone membranes for enhanced removal of natural organic matter and

alleviation of membrane fouling, Chemical Engineering Journal (2019) 122919.

[52] M.A.A. Shahmirzadi, S.S. Hosseini, G. Ruan, N. Tan, Tailoring PES nanofiltration

membranes through systematic investigations of prominent design, fabrication and operational

parameters, RSC Advances 5 (2015) 49080-49097.

[53] J. Garcia-Ivars, M.-I. Alcaina-Miranda, M.-I. Iborra-Clar, J.-A. Mendoza-Roca, L. Pastor-

Alcañiz, Enhancement in hydrophilicity of different polymer phase-inversion ultrafiltration

membranes by introducing PEG/Al2O3 nanoparticles, Separation and Purification Technology

128 (2014) 45-57.

48
[54] S. Deng, X. Liu, J. Liao, H. Lin, F. Liu, PEI modified multiwalled carbon nanotube as a

novel additive in PAN nanofiber membrane for enhanced removal of heavy metal ions, Chemical

Engineering Journal (2019) 122086.

[55] H.A. Qdais, H. Moussa, Removal of heavy metals from wastewater by membrane processes:

a comparative study, Desalination 164 (2004) 105-110.

[56] S. Hosseini, S. Amini, A. Khodabakhshi, E. Bagheripour, B. Van der Bruggen, Activated

carbon nanoparticles entrapped mixed matrix polyethersulfone based nanofiltration membrane

for sulfate and copper removal from water, Journal of the Taiwan Institute of Chemical

Engineers 82 (2018) 169-178.

[57] G.S. Ibrahim, A.M. Isloor, A.M. Asiri, A. Ismail, R. Kumar, M.I. Ahamed, Performance

intensification of the polysulfone ultrafiltration membrane by blending with copolymer

encompassing novel derivative of poly (styrene-co-maleic anhydride) for heavy metal removal

from wastewater, Chemical Engineering Journal 353 (2018) 425-435.

[58] I.H. Alsohaimi, S.M. Wabaidur, M. Kumar, M.A. Khan, Z.A. Alothman, M.A. Abdalla,

Synthesis, characterization of PMDA/TMSPEDA hybrid nano-composite and its applications as

an adsorbent for the removal of bivalent heavy metals ions, Chemical Engineering Journal 270

(2015) 9-21.

[59] E. Nightingale Jr, Phenomenological theory of ion solvation. Effective radii of hydrated

ions, The Journal of Physical Chemistry 63 (1959) 1381-1387.

[60] G. Zhou, J. Luo, C. Liu, L. Chu, J. Crittenden, Efficient heavy metal removal from

industrial melting effluent using fixed-bed process based on porous hydrogel adsorbents, Water

research 131 (2018) 246-254.

49
[61] D. Ji, C. Xiao, S. An, J. Zhao, J. Hao, K. Chen, Preparation of high-flux PSF/GO loose

nanofiltration hollow fiber membranes with dense-loose structure for treating textile wastewater,

Chemical Engineering Journal 363 (2019) 33-42.

[62] R.N. Maalige, K. Aruchamy, A. Mahto, V. Sharma, D. Deepika, D. Mondal, S.K. Nataraj,

Low operating pressure nanofiltration membrane with functionalized natural nanoclay as

antifouling and flux promoting agent, Chemical Engineering Journal 358 (2019) 821-830.

[63] R. Zhang, M. He, D. Gao, Y. Liu, M. Wu, Z. Jiao, Y. Su, Z. Jiang, Polyphenol-assisted in-

situ assembly for antifouling thin-film composite nanofiltration membranes, Journal of

membrane science 566 (2018) 258-267.

[64] K.N. Palani, N. Ramasamy, K.V. Palaniappan, Y.S. Huh, B. Natesan, Development of

integrated membrane bioreactor and numerical modeling to mitigate fouling and reduced energy

consumption in pharmaceutical wastewater treatment, Journal of Industrial and Engineering

Chemistry 76 (2019) 150-159.

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

50
Highlights

 The boehmite functionalized with curcumin was used for synthesize of membranes.

 The PES/B-Cur membranes showed high permeation flux and antifouling resistance.

 The PES/B-Cur0.5 membrane exhibited high rejection towards heavy metal ions.

 The modified membranes showed high adsorption capacity for heavy metal ions.

51

You might also like