Next Article in Journal
Follow-Up of Liver Stiffness with Shear Wave Elastography in Chronic Hepatitis C Patients in Sustained Virological Response Augments Clinical Risk Assessment
Previous Article in Journal
Analysis the Drivers of Environmental Responsibility of Chinese Auto Manufacturing Industry Based on Triple Bottom Line
Previous Article in Special Issue
Plant-Based Tacca leontopetaloides Biopolymer Flocculant (TBPF) Produced High Removal of Heavy Metal Ions at Low Dosage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pressure-Driven Membrane Process: A Review of Advanced Technique for Heavy Metals Remediation

1
Department of Chemical Engineering, Indian Institute of Technology, Ropar 140001, India
2
Department of Chemical Engineering, Indian Institute of Technology, Roorkee 247667, India
*
Authors to whom correspondence should be addressed.
Processes 2021, 9(5), 752; https://doi.org/10.3390/pr9050752
Submission received: 24 March 2021 / Revised: 21 April 2021 / Accepted: 22 April 2021 / Published: 24 April 2021
(This article belongs to the Special Issue Design and Applications of Polymeric Flocculants)

Abstract

:
Pressure-driven processes have come a long way since they were introduced. These processes, namely Ultra-Filtration (UF), Nano-Filtration (NF), and Reverse-Osmosis (RO), aim to enhance the efficiency of wastewater treatment, thereby aiming at a cleaner production. Membranes may be polymeric, ceramic, metallic, or organo-mineral, and the filtration techniques differ in pore size from dense to porous membrane. The applied pressure varies according to the method used. These are being utilized in many exciting applications in, for example, the food industry, the pharmaceutical industry, and wastewater treatment. This paper attempts to comprehensively review the principle behind the different pressure-driven membrane technologies and their use in the removal of heavy metals from wastewater. The transport mechanism has been elaborated, which helps in the predictive modeling of the membrane system. Fouling of the membrane is perhaps the only barrier to the emergence of membrane technology and its full acceptance. However, with the use of innovative techniques of fabrication, this can be overcome. This review is concluded with perspective recommendations that can be incorporated by researchers worldwide as a new problem statement for their work.

1. Introduction

The rapid industrial revolution has led to the release of heavy metals into the water streams. Even the slightest of exposure to these elements is believed to cause catastrophic consequences [1]. These elements have atomic weights of between 63.5 and 200.6 and have a specific gravity higher than 5 [2]. The life span of these elements in the ecosystem, owing to their property of bio-magnification, makes the situation worse. Therefore, heavy metal removal from industrial wastewater has become an immediate matter of concern worldwide. The most toxic heavy metals are Zinc, Chromium, Nickel, Lead, Cadmium, and Mercury. There are conventional methods to treat wastewater, such as adsorption [3], flotation [4], bio-sorption [5], coagulation [6], ion exchange [7], bioremediation [8], and electro-dialysis [9]. Although these have been well-established, restrictive environmental legislation, more extensive space requirements, labour-intensive operations, lower selectivity, lower separation efficiency, and the high cost of these conventional methods have resulted in the search for more promising and unconventional techniques. Membrane technology in industrial pollution abatement has attracted research interest worldwide due to its high efficiency, smooth operation, and less space requirements.
Membrane technology may help industrial effluents to stay within permissible standards. Treatment through membranes is highly effective and, therefore, a review of the research carried out up to now is quite imperative. There is no detailed paper that covers the application of all the pressure-driven processes in environmental applications and the implications of various parameters on the performance of the membrane process. This manuscript aims to provide a comprehensive perspective and review of the available literature. This will help researchers to identify the gaps and proceed accordingly. Figure 1 presents the comparison of various kinds of pressure-driven membrane technologies.

2. Results and Discussion

2.1. Ultrafiltration in Heavy Metal Removal

Ultrafiltration (UF) is the technique under membrane separation where the transmembrane pressure required is relatively low. This method aims to eliminate dissolved as well as colloidal particles. The major limitation of this type of pressure-driven process is the larger size of the pore. The pore size for UF lies between 2 nm and 50 nm [10]. The pore sizes are more significant than the metal ions in their hydrated forms.
Permeate fluxes are determined by varying applied pressure at a fixed temperature. The flux can be predicted by the following equation known as the Hagen- Poiseuille equation [11]:
J = P T R M + R F + R G
where J is the flux through the membrane (m3/m2.s), PT is the transmembrane pressure (N/m2), RM is the intrinsic membrane resistance (N.s/m3), RF is the fouling resistance, and RG is the resistance related to concentration polarization. RM can also be written as:
R M = 32 × Δ X × µ ε × d p 2
where dp is the mean diameter of the pore (m), µ is the viscosity of the fluid (N.S/m2), ΔX is the thickness of the membrane skin (m), and ε is the porosity of the surface of the membrane.
The performance of UF can be increased by adding micelles or polymers, which can aggregate the heavy metal particle. These are referred to as Micellar Enhanced UF (MEUF), and Polymer Enhanced UF (PEUF) (Table 1). Scamehorn first discovered MEUF for water remediation [12]. Figure 2 shows the process of micellar enhanced filtration. In the process of MEUF, the addition of surfactants either at levels equal to or higher than Critical Micelle Concentration (CMC), lead to the formation of aggregates called micelles. The solubility of metal ions in these micelles is higher and is dependent on electrostatic or Van der Waals forces. These micelles containing metal ions are then subjected to UF membrane, and thus pure water is achieved. The retention of these heavy metal bound micelles is subsequently obtained by using a UF membrane with a pore size smaller than the size of the micelle. To achieve maximum efficiency, surfactants having electrical charges are used worldwide. As an example, Sodium Dodecyl Sulphate (SDS), having a negative charge, is used to treat wastewater containing metal ions. The removal efficiency is dependent on various parameters, including the pressure which is applied, the concentration of the surfactant, the temperature of feed, the flow rate, and the concentration of the feed. Since the driving force itself is the pressure difference, it is evident that permeate flux will vary with applied pressure at a particular surfactant concentration. However, the applied pressure should be less than the maximum pressure the membrane can withstand [13]. pH also plays a vital role in the removal of heavy metals by MEUF. It was observed that the removal of Cr(VI) by MEUF using Cetyl Trimethyl Ammonium Bromide (CTAB) was maximum (90%) at a pH of 2 [14]. The surfactant to metal molar ratio is an essential variable in MEUF as the formation of micelles depends on the Critical Micelle Concentration (CMC). It has been reported that a rejection coefficient as high as 99% can be obtained by keeping the surfactant to metal molar ratio above 5 [15]. In some unusual cases, it has been observed that metal rejection coefficients as high as 90% could be achieved for Cd(II), Cu(II), Pb(II), and Zn(II) by using the concentration of surfactants less than the Critical Micelle Concentration [16]. This may be attributed to the breakage of aggregates into smaller aggregates at a much higher concentration than CMC. The removal of chromium has also been investigated by using Cetyl Pyridinium Chloride (CPC), and it was reported that the permeate concentration increased with feed concentration. The concentration was observed to increase beyond the Critical Micelle Concentration (CMC) of 43 mM [17]. The removal of Cr(VI) by using CTAB, aggregation was observed at a CTAB concentration of 0.72 mM [18]. Feed temperature is also an important parameter as the thermal expansion of the membrane as well as the viscosity of the solution is directly dependent on the temperature of the solution. Besides, Critical Micelle Concentration (CMC) is dependent on the temperature. A hybrid system containing MEUF and an Activated Carbon Filter (ACF) was employed to obtain 96.2% removal efficiency of Ni(II) [19].
Polymer Enhanced UF (PEUF) is also a plausible solution to wastewater treatment. (Table 2). In this method, a water-soluble polymer is added as a complexing agent with metal ions and forms a macromolecule. The molecular weight of the macromolecule is higher than the molecular weight cut off of the membrane. Therefore, the metallic ions, along with the complexing agent, are retained over the UF membrane. The most commonly used complexing agents are Polyacrylic acid (PAA) [39], Polyvinylamine (PVA) [40], Polyethyleneimine (PEI) [41], a copolymer of maleic acid and acrylic acid (PMA) [42], and humic acid [43]. The performance of this method is dependent on the type of metal to be removed and the type of polymer used. Besides, solution pH and the existence of other ions in the solution also affect the performance. The efficiency of this method is dependent mainly on the type of metal and polymer, the feed concentration, the pH of the solution, and the presence of other salts. It was reported that the rejection coefficient reduced from 90% to 32% when the Cr(VI) concentration in the feed was increased from 25 PPM to 400 PPM [44]. The optimum polymer/metal weight ratio for the selective removal of Ni(II) and Cu(II) by using complexing agent PEI were 6 and 3, respectively [45]. The polymer/metal weight ratio of 25 was observed to be the most suitable for the removal of Ni(II) and Cr(III) with removal efficiencies of 99.5% and 99.8%, respectively [46]. The solution pH also affects the performance of the system. The complete removal of Cr(VI) by PEUF using Poly(N,N dimethylaminoethyl methacrylate) was reported when the pH was kept at 4 [47].

2.2. Nanofiltration for the Removal of Heavy Metals from Wastewater

Nanofiltration (NF) possesses properties between those of UF and Reverse osmosis (RO), and therefore the pore size is usually less than 2 nm, corresponding to an MWCO of 100–1000 Da [63]. The presence of surface functional groups and their dissociation provides charge to these membranes. NF membranes’ charge is highly dependent on solution pH. For example, typical polyamide NF membranes have an isoelectric point (IEP) between 3.5–5. At pH lower than IEP, amine groups and carboxylic groups are protonated (R-NH2+/RCOOH), which confer to the membrane a positive charge. Contrarily, at pH>IEP, NF membranes exhibit a negative charge because of the deprotonation of the above-mentioned functional groups (R-NH/R-COO). In view of the fact that most of the heavy metals are cations, NF is of prime importance due to the separation achieved by the combination of steric effect and electrostatic forces [64]. Higher rejection of divalent ions, lower rejection of monovalent ions, and higher flux compared to RO membranes are some of the critical attributes of NF membranes. To reduce energy costs, NF is extensively used in the wastewater treatment processes and is currently trying to replace RO to make the processes more economically viable. The NF membranes separate the solute from the solution via two mechanisms. The first is known as ionic separation and corresponds to the separation based on the charge of solute in water. The second is known as sieving, which corresponds to the molecular weight of uncharged solutes. The non-sieving rejection mechanisms of NF are Donnan exclusion [65] and dielectric exclusion [66]. This is being explored to remove heavy metals from wastewater (Table 2) due to its ease of operation, reliability, and lower energy consumptions. The NF technique has been applied for removal of heavy metals such as Copper [67], Cobalt [68], Nickel [69], Zinc [70], Lead [71], Cadmium [72], Chromium [73], Arsenic [74] and mercury [75].
It was reported that As(V) removal increased with increased pH. The rejection increased from 74% to 88% when the pH was increased from 3.4 to 10. This could be explained by the fact that the monovalent ion H2AsO4 is dominant in the range of pH 4–6 while the divalent ion HAsO42− is dominant above pH 7. Owing to the large, hydrated radii of divalent ions, they are rejected by the membrane at a much higher rate. The As(V) removal increased with lower temperature. It was observed that at 15 °C, the arsenic removal was 95.4% which decreased to 93.1% on increasing the temperature to 40 °C. This is because of the increased diffusivity of arsenic with temperature [76]. In another study, it was observed that natural organic matter, humic acid, increased the rejection coefficient of As(V) by using NF membranes. The rejection coefficient for all four types of membranes was recorded to be higher than 94% [74]. Another study reported that the removal of Ni(II) ions increased with feed pressure and feed concentration. The maximum observed rejection coefficient of Ni(II) was 98.94% for an initial feed concentration of 5 PPM [77]. The efficiency of silver recovery was found to be 29%–59% for hybrid cyanidation and membrane separation [78]. Modification of NF membranes at the laboratory scale has also been done to increase their selectivity. A study reported removal efficiencies of 47.9, 44.2, 52.3 for Cu(II), Cd(II) and Cr(VI), respectively by the NF membrane modified by halloysite nanotubes (HNTs) functionalized with 3-aminopropyltriethoxysilane (APTES) [79]. Polyether sulphone (PES) NF membranes were grafted with poly- (amidoamine) dendrimer and showed outstanding performance with almost 99% removal efficiency for Pb(II), Cu(II), Ni(II), Cd(II), Zn(II), and As(V) [80]. Chelating polymers like poly- (acrylic acid-co-maleic acid) (PAM) and poly- (acrylic acid) (PAA) were adsorbed on the NF filtration membranes to treat Cd(II), Zn(II), Pb(II), Ni(II), Cu(II), As(V), and Cr(VI), and rejection coefficients of more than 98% were obtained for all the heavy metals [81]. Hollow fibre NF membranes have been fabricated using polybenzimidazole (PBI) and were investigated for their Cr(VI) removal capacity. The fabricated PBI membranes with Molecular Weight Cut Off (MWCO) of 525 Da showed 95.7% Cr(VI) removal [82]. The removal of Cr(VI) was also attempted by fabricating asymmetric NF membrane with poly- (m-phenylene isophthalamide) (PMIA). It was observed that 98% of Cr(VI) was removed at a pH of 8 with the fabricated membrane [83]. The presence of metal complexing polymer also played a crucial role in the application of membranes in the treatment of heavy metal. It was seen that with the addition of Bovine Serum Albumin (BSA), rejection coefficients of 93%, 99%, 93%, and 99% were obtained by employing a ceramic membrane with MWCO of 450 Da for Zn(II), Cd(II), Pb(II) and Cu(II), respectively [84]. The capillary UF properties were merged with capillary NF to achieve the desired separation of heavy metal from wastewater. This combined technique, called Direct Capillary NF (CNF), was employed to remove Pb(II), and 83% removal efficiency was obtained with the flux of 20 L/m2·h [85]. NF was also combined with MF, and the hybrid system was able to obtain 99% removal efficiency of Cr(VI) from aqueous solutions [86]. A dual-layer NF membrane was fabricated with an outer layer of polybenzimidazole (PBI) and a blend of polyethersulfone (PES) and polyvinylpyrrolidone (PVP). This membrane with an active area of 0.0037 m2 was employed to remove Cr(VI), Pb(II), and Cd(II). It was observed that 98% of Cr(VI) and 93% of Pb(II) were removed at pH of 12 and 2.2, respectively [87]. One study showed the laboratory scale development of an amphoteric hollow fiber NF membrane which was capable of removing 97% of As(V)) at a pH of 10 [88]. A positively charged membrane was fabricated by modifying NF membrane with hyperbranched polyethyleneimine (PEI) and removed 91.05% of Pb(II) at a pH of less than 8 [89]. A positively charged NF membrane was developed with the use of 2-chloro-1-methyliodopyridine and removed 96% and 95.8% of Cu(II) and Ni(II), respectively [90]. Nanotechnology has also come a long way since its development. Some researchers blended 0.5 weight% of cobalt ferrite nanoparticles in a polyethersulphone NF membrane and achieved rejection of 98%, 92%, and 88% for Cu(II), Ni(II) and Pb(II), respectively [91]. Beside, 1 wt% of cellulose nanocrystals functionalized by amine groups were also embedded in the PES membrane and the fabricated membrane removed 90% of Cu(II) ions [92]. A NF membrane Desal 5 DK was investigated for separating Cr(III) ions from acid solutions. It was evaluated that the Cr(III) concentration profile across the membrane is sensitive to the initial concentration of chromium as well as to the pore dielectric constant [93]. The performance of a semi-aromatic poly(piperazineamide) membrane was assessed for the removal of metals from copper metallurgical process streams. The membrane showed high metal rejections of more than 80%, and the high rejection values were associated with dielectric exclusion [94]. A research group studied the performance of extreme-acid-resistant duracid membrane for the valorisation of copper acidic effluent. The permeate flux and the feed water composition were varied, and it was observed that the metal rejection was more than 90% [95]. A novel nanocomposite membrane was developed for the removal of heavy metals, and its performance was compared to the polymeric membranes. It was concluded that while the polymeric membranes can exhibit a rejection from 77 up to 99%, the nanocomposite membranes can completely reject heavy metals (up to 100%) [96]. Table 3 summarizes the recent literature of various NF membranes used for heavy metal removal and their performance.

2.3. Reverse Osmosis in Heavy Metal Removal from Wastewater

The technique of reverse osmosis (RO) is the most efficient in terms of the contaminants it can separate from water. The semi-permeable membrane mostly allows water to pass and retains most of the pollutants. This technique accounts for more than 20% of the world’s desalination capacity [97]. The application of RO in the removal of heavy metals from wastewater is being investigated (Table 4). Low-pressure RO has been applied with a complexing agent to remove Ni(II) and Cu(II), and the removal efficiencies for both single and a mixture of ions achieved over 99% [98]. The complete removal of Ni(II) was obtained by employing the UF/RO hybrid system in the metal finishing industry [99]. A rejection coefficient as high as 99.9% was achieved for removal of Cr(VI) by employing a RO membrane at a pH of 8 [100]. The pH of the solution plays a crucial role in the removal of heavy metals by RO. Researchers obtained 91% removal of Cr(VI) at a pH of 3 [101]. For an initial feed concentration of 200 mg/L, 99% of Cd(II) was rejected by using RO from contaminated wastewater [102]. Commercially available polyamide ultra-low pressure reverse osmosis (ULPRO) and a NF membrane were employed to separate heavy metals. The rejection of heavy metals was achieved to be greater than 97% [103]. A study of the implementation of sequential stages of MF, NF and RO to separate noble metals from a gold mining effluent was carried out. It was observed that the retention of metals in the concentrates of MF and RO was above 95% [104]. Recovery of heavy metals such as Mn, Fe, Cu, Zn, As, Cd, and Pb was assessed using a volume retarded osmosis low-pressure membrane (VRO-LPM), and 95% rejection was obtained for all the heavy metals [105].
Table 3. Review of literature concerning application of NF in heavy metal removal.
Table 3. Review of literature concerning application of NF in heavy metal removal.
Characteristics of MembraneHeavy Metal TargetedInitial Metal Concentration (mg/L)pHPressure (bar)Removal Efficiency (%)Ref
Flat organic membranes
MWCO = 1000 Da
Permeability = 1.6 × 10−3 L m−2 s−1 bar−1
Ni(II)
Cu(II)
691.21476.1
72.6
[106]
Polyamide membrane supported by diaminobenzenesulfonic acid (DABSA)
MWCO = 500 Da
Permeability = 11.8 L m−2 h−1 bar−1
Negative charged
Cr(VI)46094 >99[107]
Chitosan/polyvinyl alcohol/montmorillonite clay membrane
Active area = 0.00385 m2
Cr(VI)5071(84–88.34)[108]
PAN/Sulfonated Polyarylene ether benzonitrile
MWCO = 300 Da
Permeability = 7.62 LMH.bar−1
Pb(II)
Cd(II)
1000(2–5)694.6
95.1
[109]
Spiral wound Polymeric membrane (NF270–2540)
Active area = 2.6 m2
Negatively charged
MWCO = (200–400) Da
Co(II)
Ni(II)
20(3.4–5.6)
3.4
6
10
100
91.94
[110]
Polyamide membrane (NF270)
Active area = 0.00076 m2.
The surface of the membrane is positively charged for pH less than 4 and negatively charged for a higher value.
Cd(II)
Mn(II)
Pb(II)
10001.5499
89
74
[111]
Polyamide membrane
Active area = 0.47 m2
Isoelectric point is at pH 3.3–4
Pb(II)
Ni(II)
15.56–886
93
[112]
Polyethersulfone Membrane
Area = 0.00001256 m2
Membrane fabricated with 0.5 wt% magnetic graphene based nanocomposites
Cu(II)205492[113]
Thin-film composite.
Area = 0.024 m2
The isoelectric point is 3.6.
Membrane surface is negatively charged
Pb(II)1505.82599[114]
Polyamide thin film composite membrane
MWCO = 400
Zeta potential = −36.8 mV
Cr(VI)
As(V)
0.188.18(25–95)[115]
NF 300
Active area = 2.5 m2
As(V)0.015 to 0.0255099.8[116]
Aromatic polyamide membrane
MWCO = (200–300)
Water permeability = 2.4 × 10−8 m3/(s.m2.kPa)
As(V)0.27.510>94[74]
Thin film composite membrane (NF 2)
Active area = 0.01 m2
As(V)0.150–0.252711.76(97–100)[117]
NF Hollow fibre membrane
MWCO = 520 Da
Isoelectric point = 6.6
Pure water flux = 47.5 L/(m2.h)
Ni(II)
Cr(VI)
Cu(II)
142.23
121.23
56.55
2.31494.99
95.76
95.33
[118]
Polyamide composite membrane
MWCO ranges between 150 and 300
Effective area = 0.00572 m2
Cu(II)2304.56.8998.1[119]
Thin-film Composite membrane
Area = 0.0036 m2
Pb(II)400 330 97.5[120]
Thin-film composite membrane (AFC 40)
Effective area = 0.024 m2
Co(II)100 325 97[64]
NF 300 membrane
Effective area = 0.015 m2
Cd(II)
Ni(II)
5 52097.26
98.90
[121]
Polyamide membrane (NF270)
Membrane surface area = 0.0012 m2
Isoelectric point = 3.3
Cu(II)25,000(3–10)3099.5[122]
Negatively charged microporous NF, Nanomax50Cu(II)200 < 4.53 66[123]
NF spiral-wound membraneCu(II)50 53.8 >95[124]
Polyamide membrane (NTR 729HF)
MWCO = 700
pH 6.5
Effective membrane area = 0.006 m2
As(V)
As(III)
0.5(5–9)0.1–581
57
[125]
Polyamide thin film composite (NF90- 2540)
Active surface area = 2.6 m2
MWCO = 200 Da
As(V)0.1 86>90[76]
Composite polyamide spiral wound membrane(NFI)
Membrane area = 0.75 m2
Pure water permeability = 3.20 L/hm2 bar
Cr(VI)1000 5–8 99[73]
Table 4. Review of literature concerning application of RO in heavy metal removal.
Table 4. Review of literature concerning application of RO in heavy metal removal.
Characteristics of MembraneHeavy Metal TargetedInitial Metal Concentration (mg/L)Operating ConditionsPressure (bar)Removal EfficiencyRef
TFC spiral wound membrane
Active area = 1.95 m2
Allowable operating pH range = 4–11
Max operating temperature = 45 °C
Max feed turbidity, NTU = 1
Max feed SDI = 5
Cu(II)
Ni(II)
500 Na2EDTA was added as a chelating agent at pH 55.0699.5[126]
Disk membranes
Polyamide selective layer is supported on the polysulfone layer
Cu(II)20–100 Addition of Sodium dodecyl sulphate increased the removal efficiencyLow pressure 70–95[127]
Polyamide thin film composite membrane
MWCO = N.A
Pure water permeability = 0.75 L m−2 d−1 kPa−1
pH range = 3–10
Zeta potential = −4.5 mV
Cr(VI)
As(V)
0.1 85.13>90[106]
Polyamide membrane
Active area = 0.014 m2
As(V)0.1 10–4099.75[128]
Brackish water membrane
Active surface area = 0.014 m2
As(III) pH 9.6 4099[129]
SWHR membrane (Filmtec)As(V)
As(III)
0.2pH 4
pH 9.1
10–3596.8
92.5
[130]
Polyamide spiral wound membrane
Membrane surface area = 2.5 m2
pH range = (4–11)
Cu(II)
Cd(II)
500 1398
99
[131]
SE and MPF44 NF membranes
Active membrane surface area = 0.0028 m2
Cu(II)2 M 35>95[67]

2.4. Fouling of Membranes

Despite its potential in the treatment of wastewater, there are certain limitations that prohibit the application of membranes in large-scale operations. The most challenging of all limitations arise from membrane fouling caused by different inorganic salts, which decreases permeate flux and increases feed pressure. The quality of the product is bad, and the life of the membrane is shortened [132]. The impacts of various model foulants on the performance of RO and NF membranes has also been studied. It was observed that organic foulants such as humic acid and sodium alginate caused the most severe drop in permeate flux [133]. The research was conducted to evaluate the cost of fouling in full-scale RO and NF installations, and it was found that the cost of fouling in RO plants was around 24% of operational expenses. The cost for early membrane replacement accounted for the significant portion of the total cost. It was concluded that cleaning-in-place automation could save up to 3% of operational expenses [134]. A surface-patterned alumina ceramic membrane with a porous gradient structure was fabricated to improve the anti-fouling ability [135]. The effect of gradient profile in ceramic membranes on fouling was also studied by a research group, and it was found that gradient profile reduced membrane resistance and fouling [136]. A hydrogenated TiO2 membrane with photocatalytically enhanced anti-fouling was developed for the treatment of surface water. It was observed that upon UV irradiation, the hydrogenated TiO2 membrane showed 60% higher humic acid removal efficiency as compared to the pristine TiO2 membrane [137].

3. Conclusions

Membrane processes are the best available techniques for water and wastewater treatment. The capability of separating a wide variety of components from the aqueous stream makes membrane technology one of the most promising methods available. A great deal of research has been done on heavy metal removal by ultra-filtration using surfactants and complexing agents. There is no literature available on separation by using a hybrid system of MEUF and PEUF for heavy metal removal. Although work has been done on the application of nanotechnology in membrane processes, little has been done on their application in the removal of heavy metals from wastewater. The nano-filtration membranes have been modified by using specific materials to enhance their selectivity. The combination of nanoparticles and nanocomposites with membranes seems to be a research gap. Besides, the fouling of the membrane is a significant limitation in the application of membranes in wastewater treatment. New fabrication techniques of membrane development, such that fouling of the membranes can be controlled as well as mitigated, are highly recommended.

Author Contributions

Conceptualization, B.V. and C.B.; methodology, B.V.; software, B.V.; validation, B.V. and C.B.; formal analysis, B.V.; investigation, B.V.; resources, C.B.; data curation, B.V.; writing—original draft preparation, B.V.; writing—review and editing, S.P.G. and M.S.; visualization, B.V.; supervision, C.B., S.P.G., and M.S.; project administration, C.B.; funding acquisition, C.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bolisetty, S.; Peydayesh, M.; Mezzenga, R. Sustainable technologies for water purification from heavy metals: Review and analysis. Chem. Soc. Rev. 2019, 48, 463–487. [Google Scholar] [CrossRef]
  2. Ihsanullah; Abbas, A.; Al-Amer, A.M.; Laoui, T.; Al-Marri, M.J.; Nasser, M.S.; Khraisheh, M.; Atieh, M.A. Heavy metal removal from aqueous solution by advanced carbon nanotubes: Critical review of adsorption applications. Sep. Purif. Technol. 2016, 157, 141–161. [Google Scholar] [CrossRef]
  3. Hong, M.; Yu, L.; Wang, Y.; Zhang, J.; Chen, Z.; Dong, L.; Zan, Q.; Li, R. Heavy metal adsorption with zeolites: The role of hierarchical pore architecture. Chem. Eng. J. 2019, 359, 363–372. [Google Scholar] [CrossRef]
  4. Taseidifar, M.; Ziaee, M.; Pashley, R.M.; Ninham, B.W. Ion flotation removal of a range of contaminant ions from drinking water. J. Environ. Chem. Eng. 2019, 7, 103263. [Google Scholar] [CrossRef]
  5. Jaafari, J.; Yaghmaeian, K. Optimization of heavy metal biosorption onto freshwater algae (Chlorella coloniales) using response surface methodology (RSM). Chemosphere 2019, 217, 447–455. [Google Scholar] [CrossRef]
  6. Ahmed, M.J.K.; Ahmaruzzaman, M. A review on potential usage of industrial waste materials for binding heavy metal ions from aqueous solutions. J. Water Process Eng. 2016, 10, 39–47. [Google Scholar] [CrossRef]
  7. Patil, D.S.; Chavan, S.M.; Oubagaranadin, J.U.K. A review of technologies for manganese removal from wastewaters. J. Environ. Chem. Eng. 2016, 4, 468–487. [Google Scholar] [CrossRef]
  8. Aslam, A.; Thomas-Hall, S.R.; Mughal, T.; Zaman, Q.U.; Ehsan, N.; Javied, S.; Schenk, P.M. Heavy metal bioremediation of coal-fired flue gas using microalgae under different CO2 concentrations. J. Environ. Manag. 2019, 241, 243–250. [Google Scholar] [CrossRef] [PubMed]
  9. Kirkelund, G.M.; Jensen, P.E.; Ottosen, L.M.; Pedersen, K.B. Comparison of two- and three-compartment cells for electrodialytic removal of heavy metals from contaminated material suspensions. J. Hazard. Mater. 2019, 367, 68–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Pendergast, M.T.; Hoek, E.M.V. A review of water treatment membrane nanotechnologies. Energy Environ. Sci. 2011, 4, 1946–1971. [Google Scholar] [CrossRef] [Green Version]
  11. Yang, J.S.; Baek, K.; Yang, J.W. Crossflow ultrafiltration of surfactant solutions. Desalination 2005, 184, 385–394. [Google Scholar] [CrossRef]
  12. Mungray, A.A.; Kulkarni, S.V.; Mungray, A.K. Removal of heavy metals from wastewater using micellar enhanced ultrafiltration technique: A review. Cent. Eur. J. Chem. 2012, 10, 27–46. [Google Scholar] [CrossRef]
  13. Chhatre, A.J.; Marathe, K.V. Dynamic analysis and optimization of surfactant dosage in micellar enhanced ultrafiltration of nickel from aqueous streams. Sep. Sci. Technol. 2006, 41, 2755–2770. [Google Scholar] [CrossRef]
  14. Nura, C.S.; Chattree, A.; Singh, R.P.; Nath, S. Removal of hexavalent chromium by mixed micelles of cetyl trimethyl ammonium bromide using micellar enhanced ultrafiltration. IJCS 2017, 5, 627–631. [Google Scholar]
  15. Landaburu-Aguirre, J.; García, V.; Pongrácz, E.; Keiski, R.L. The removal of zinc from synthetic wastewaters by micellar-enhanced ultrafiltration: Statistical design of experiments. Desalination 2009, 240, 262–269. [Google Scholar] [CrossRef]
  16. Samper, E.; Rodríguez, M.; De la Rubia, M.A.; Prats, D. Removal of metal ions at low concentration by micellar-enhanced ultrafiltration (MEUF) using sodium dodecyl sulfate (SDS) and linear alkylbenzene sulfonate (LAS). Sep. Purif. Technol. 2009, 65, 337–342. [Google Scholar] [CrossRef]
  17. Ghosh, G.; Bhattacharya, P.K. Hexavalent chromium ion removal through micellar enhanced ultrafiltration. Chem. Eng. J. 2006, 119, 45–53. [Google Scholar] [CrossRef]
  18. Nguyen, H.T.; Chang, W.S.; Nguyen, N.C.; Chen, S.S.; Chang, H.M. Influence of micelle properties on micellar-Enhanced ultrafiltration for chromium recovery. Water Sci. Technol. 2015, 72, 2045–2051. [Google Scholar] [CrossRef]
  19. Rafique, R.F.; Lee, S. Micellar Enhanced Ultrafiltration (MEUF) and activated carbon fiber (ACF) hybrid processes for the removal of cadmium from an aqueous solution. Korean Chem. Eng. Res. 2014, 52, 775–780. [Google Scholar] [CrossRef] [Green Version]
  20. Tortora, F.; Innocenzi, V.; Prisciandaro, M.; Mazziotti di Celso, G.; Vegliò, F. Analysis of membrane performance in Ni and Co removal from liquid wastes by means of micellar-enhanced ultrafiltration. Desalin. Water Treat. 2016, 57, 22860–22867. [Google Scholar] [CrossRef]
  21. Bahmani, P.; Maleki, A.; Rezaee, R.; Khamforosh, M.; Yetilmezsoy, K.; Dehestani Athar, S.; Gharibi, F. Simultaneous removal of arsenate and nitrate from aqueous solutions using micellar-enhanced ultrafiltration process. J. Water Process Eng. 2019, 27, 24–31. [Google Scholar] [CrossRef]
  22. Huang, J.H.; Zeng, G.M.; Zhou, C.F.; Li, X.; Shi, L.J.; He, S.B. Adsorption of surfactant micelles and Cd2+/Zn2+ in micellar-enhanced ultrafiltration. J. Hazard. Mater. 2010, 183, 287–293. [Google Scholar] [CrossRef] [PubMed]
  23. Landaburu-Aguirre, J.; Pongrácz, E.; Perämäki, P.; Keiski, R.L. Micellar-enhanced ultrafiltration for the removal of cadmium and zinc: Use of response surface methodology to improve understanding of process performance and optimisation. J. Hazard. Mater. 2010, 180, 524–534. [Google Scholar] [CrossRef] [PubMed]
  24. Danis, U.; Aydiner, C. Investigation of process performance and fouling mechanisms in micellar-enhanced ultrafiltration of nickel-contaminated waters. J. Hazard. Mater. 2009, 162, 577–587. [Google Scholar] [CrossRef]
  25. Huang, J.; Yuan, F.; Zeng, G.; Li, X.; Gu, Y.; Shi, L.; Liu, W.; Shi, Y. Influence of pH on heavy metal speciation and removal from wastewater using micellar-enhanced ultrafiltration. Chemosphere 2017, 173, 199–206. [Google Scholar] [CrossRef]
  26. Verma, S.P.; Sarkar, B. Simultaneous removal of Cd (II) and p-cresol from wastewater by micellar-enhanced ultrafiltration using rhamnolipid: Flux decline, adsorption kinetics and isotherm studies. J. Environ. Manag. 2018, 213, 217–235. [Google Scholar] [CrossRef]
  27. Innocenzi, V.; Prisciandaro, M.; Tortora, F.; Mazziotti di Celso, G.; Vegliò, F. Treatment of WEEE industrial wastewaters: Removal of yttrium and zinc by means of micellar enhanced ultra filtration. Waste Manag. 2018, 74, 393–403. [Google Scholar] [CrossRef]
  28. Lee, S.H.; Shrestha, S. Application of micellar enhanced ultrafiltration (MEUF) process for zinc (II) removal in synthetic wastewater: Kinetics and two-parameter isotherm models. Int. Biodeterior. Biodegrad. 2014, 95, 241–250. [Google Scholar] [CrossRef]
  29. Häyrynen, P.; Landaburu-Aguirre, J.; Pongrácz, E.; Keiski, R.L. Study of permeate flux in micellar-enhanced ultrafiltration on a semi-pilot scale: Simultaneous removal of heavy metals from phosphorous rich real wastewaters. Sep. Purif. Technol. 2012, 93, 59–66. [Google Scholar] [CrossRef]
  30. Abbasi-Garravand, E.; Mulligan, C.N. Using micellar enhanced ultrafiltration and reduction techniques for removal of Cr(VI) and Cr(III) from water. Sep. Purif. Technol. 2014, 132, 505–512. [Google Scholar] [CrossRef]
  31. Schwarze, M.; Groß, M.; Moritz, M.; Buchner, G.; Kapitzki, L.; Chiappisi, L.; Gradzielski, M. Micellar enhanced ultrafiltration (MEUF) of metal cations with oleylethoxycarboxylate. J. Membr. Sci. 2015, 478, 140–147. [Google Scholar] [CrossRef]
  32. Tanhaei, B.; Pourafshari Chenar, M.; Saghatoleslami, N.; Hesampour, M.; Laakso, T.; Kallioinen, M.; Sillanpää, M.; Mänttäri, M. Simultaneous removal of aniline and nickel from water by micellar-enhanced ultrafiltration with different molecular weight cut-off membranes. Sep. Purif. Technol. 2014, 124, 26–35. [Google Scholar] [CrossRef]
  33. Channarong, B.; Lee, S.H.; Bade, R.; Shipin, O.V. Simultaneous removal of nickel and zinc from aqueous solution by micellar-enhanced ultrafiltration and activated carbon fiber hybrid process. Desalination 2010, 262, 221–227. [Google Scholar] [CrossRef]
  34. Tanhaei, B.; Pourafshari Chenar, M.; Saghatoleslami, N.; Hesampour, M.; Kallioinen, M.; Sillanpää, M.; Mänttäri, M. Removal of nickel ions from aqueous solution by micellar-enhanced ultrafiltration, using mixed anionic-non-ionic surfactants. Sep. Purif. Technol. 2014, 138, 169–176. [Google Scholar] [CrossRef]
  35. Li, X.; Zeng, G.M.; Huang, J.H.; Zhang, D.M.; Shi, L.J.; He, S.B.; Ruan, M. Simultaneous removal of cadmium ions and phenol with MEUF using SDS and mixed surfactants. Desalination 2011, 276, 136–141. [Google Scholar] [CrossRef]
  36. Huang, J.; Shi, L.; Zeng, G.; Li, H.; Huang, H.; Gu, Y.; Shi, Y.; Yi, K.; Li, X. Removal of Cd(Ⅱ) by micellar enhanced ultrafiltration: Role of SDS behaviors on membrane with low concentration. J. Clean. Prod. 2019, 209, 53–61. [Google Scholar] [CrossRef]
  37. Huang, J.; Li, H.; Zeng, G.; Shi, L.; Gu, Y.; Shi, Y.; Tang, B.; Li, X. Removal of Cd(II) by MEUF-FF with anionic-nonionic mixture at low concentration. Sep. Purif. Technol. 2018, 207, 199–205. [Google Scholar] [CrossRef]
  38. Li, C.W.; Liu, C.K.; Yen, W.S. Micellar-enhanced ultrafiltration (MEUF) with mixed surfactants for removing Cu(II) ions. Chemosphere 2006, 63, 353–358. [Google Scholar] [CrossRef] [PubMed]
  39. Ennigrou, D.J.; Sik Ali, M.B.; Dhahbi, M.; Ferid, M. Removal of heavy metals from aqueous solution by polyacrylic acid enhanced ultrafiltration. Desalin. Water Treat. 2015, 56, 2682–2688. [Google Scholar] [CrossRef]
  40. Huang, Y.; Wu, D.; Wang, X.; Huang, W.; Lawless, D.; Feng, X. Removal of heavy metals from water using polyvinylamine by polymer-enhanced ultrafiltration and flocculation. Sep. Purif. Technol. 2016, 158, 124–136. [Google Scholar] [CrossRef]
  41. Aroua, M.K.; Zuki, F.M.; Sulaiman, N.M. Removal of chromium ions from aqueous solutions by polymer-enhanced ultrafiltration. J. Hazard. Mater. 2007, 147, 752–758. [Google Scholar] [CrossRef] [PubMed]
  42. Qiu, Y.R.; Mao, L.J. Removal of heavy metal ions from aqueous solution by ultrafiltration assisted with copolymer of maleic acid and acrylic acid. Desalination 2013, 329, 78–85. [Google Scholar] [CrossRef]
  43. Kim, H.J.; Baek, K.; Kim, B.K.; Yang, J.W. Humic substance-enhanced ultrafiltration for removal of cobalt. J. Hazard. Mater. 2005, 122, 31–36. [Google Scholar] [CrossRef]
  44. Muthumareeswaran, M.R.; Alhoshan, M.; Agarwal, G.P. Ultrafiltration membrane for effective removal of chromium ions from potable water. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  45. Molinari, R.; Poerio, T.; Argurio, P. Selective separation of copper(II) and nickel(II) from aqueous media using the complexation-ultrafiltration process. Chemosphere 2008, 70, 341–348. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, Q.; Gao, J.; Qiu, Y.R. Removal of Ni (II) and Cr (III) by complexation-ultrafiltration using rotating disk membrane and the selective separation by shear induced dissociation. Chem. Eng. Process. Process Intensif. 2019, 135, 236–244. [Google Scholar] [CrossRef]
  47. Sánchez, J.; Espinosa, C.; Pooch, F.; Tenhu, H.; Pizarro, G.D.C.; Oyarzún, D.P. Poly(N,N-dimethylaminoethyl methacrylate) for removing chromium (VI) through polymer-enhanced ultrafiltration technique. React. Funct. Polym. 2018, 127, 67–73. [Google Scholar] [CrossRef] [Green Version]
  48. Camarillo, R.; Pérez, Á.; Cañizares, P.; de Lucas, A. Removal of heavy metal ions by polymer enhanced ultrafiltration. Batch process modeling and thermodynamics of complexation reactions. Desalination 2012, 286, 193–199. [Google Scholar] [CrossRef]
  49. Labanda, J.; Khaidar, M.S.; Llorens, J. Feasibility study on the recovery of chromium (III) by polymer enhanced ultrafiltration. Desalination 2009, 249, 577–581. [Google Scholar] [CrossRef]
  50. Camarillo, R.; Llanos, J.; García-Fernández, L.; Pérez, Á.; Cañizares, P. Treatment of copper (II)-loaded aqueous nitrate solutions by polymer enhanced ultrafiltration and electrodeposition. Sep. Purif. Technol. 2010, 70, 320–328. [Google Scholar] [CrossRef]
  51. Llanos, J.; Camarillo, R.; Pérez, Á.; Cañizares, P. Polymer supported ultrafiltration as a technique for selective heavy metal separation and complex formation constants prediction. Sep. Purif. Technol. 2010, 73, 126–134. [Google Scholar] [CrossRef]
  52. Jellouli Ennigrou, D.; Gzara, L.; Ramzi Ben Romdhane, M.; Dhahbi, M. Cadmium removal from aqueous solutions by polyelectrolyte enhanced ultrafiltration. Desalination 2009, 246, 363–369. [Google Scholar] [CrossRef]
  53. Ennigrou, D.J.; Ben Sik Ali, M.; Dhahbi, M. Copper and Zinc removal from aqueous solutions by polyacrylic acid assisted-ultrafiltration. Desalination 2014, 343, 82–87. [Google Scholar] [CrossRef]
  54. Lam, B.; Déon, S.; Morin-Crini, N.; Crini, G.; Fievet, P. Polymer-enhanced ultrafiltration for heavy metal removal: Influence of chitosan and carboxymethyl cellulose on filtration performances. J. Clean. Prod. 2018, 171, 927–933. [Google Scholar] [CrossRef]
  55. Huang, Y.; Du, J.; Zhang, Y.; Lawless, D.; Feng, X. Batch process of polymer-enhanced ultrafiltration to recover mercury (II) from wastewater. J. Membr. Sci. 2016, 514, 229–240. [Google Scholar] [CrossRef]
  56. Chou, Y.H.; Choo, K.H.; Chen, S.S.; Yu, J.H.; Peng, C.Y.; Li, C.W. Copper recovery via polyelectrolyte enhanced ultrafiltration followed by dithionite based chemical reduction: Effects of solution pH and polyelectrolyte type. Sep. Purif. Technol. 2018, 198, 113–120. [Google Scholar] [CrossRef]
  57. Ennigrou, D.J.; Gzara, L.; Romdhane, M.R.B.; Dhahbi, M. Retention of cadmium ions from aqueous solutions by poly(ammonium acrylate) enhanced ultrafiltration. Chem. Eng. J. 2009, 155, 138–143. [Google Scholar] [CrossRef]
  58. Cañizares, P.; Pérez, Á.; Llanos, J.; Rubio, G. Preliminary design and optimisation of a PEUF process for Cr(VI) removal. Desalination 2008, 223, 229–237. [Google Scholar] [CrossRef]
  59. Barakat, M.A.; Schmidt, E. Polymer-enhanced ultrafiltration process for heavy metals removal from industrial wastewater. Desalination 2010, 256, 90–93. [Google Scholar] [CrossRef]
  60. Huang, Y.; Du, J.R.; Zhang, Y.; Lawless, D.; Feng, X. Removal of mercury (II) from wastewater by polyvinylamine-enhanced ultrafiltration. Sep. Purif. Technol. 2015, 154, 1–10. [Google Scholar] [CrossRef]
  61. Llanos, J.; Pérez, Á.; Cañizares, P. Copper recovery by polymer enhanced ultrafiltration (PEUF) and electrochemical regeneration. J. Membr. Sci. 2008, 323, 28–36. [Google Scholar] [CrossRef]
  62. Cañizares, P.; Pérez, Á.; Camarillo, R.; Llanos, J.; López, M.L. Selective separation of Pb from hard water by a semi-continuous polymer-enhanced ultrafiltration process (PEUF). Desalination 2007, 206, 602–613. [Google Scholar] [CrossRef]
  63. Tajuddin, M.H.; Yusof, N.; Wan Azelee, I.; Wan Salleh, W.N.; Ismail, A.F.; Jaafar, J.; Aziz, F.; Nagai, K.; Razali, N.F. Development of copper-aluminum layered double hydroxide in thin film nanocomposite nanofiltration membrane for water purification process. Front. Chem. 2019, 7, 1–11. [Google Scholar] [CrossRef] [Green Version]
  64. Ye, C.C.; An, Q.F.; Wu, J.K.; Zhao, F.Y.; Zheng, P.Y.; Wang, N.X. Nanofiltration membranes consisting of quaternized polyelectrolyte complex nanoparticles for heavy metal removal. Chem. Eng. J. 2019, 359, 994–1005. [Google Scholar] [CrossRef]
  65. Yaroshchuk, A.E. Non-steric mechanisms of nanofiltration: Superposition of Donnan and dielectric exclusion. Sep. Purif. Technol. 2001, 22–23, 143–158. [Google Scholar] [CrossRef]
  66. Yaroshchuk, A.E. Dielectric exclusion of ions from membranes. Adv. Colloid Interface Sci. 2000, 85, 193–230. [Google Scholar] [CrossRef]
  67. Cséfalvay, E.; Pauer, V.; Mizsey, P. Recovery of copper from process waters by nanofiltration and reverse osmosis. Desalination 2009, 240, 132–142. [Google Scholar] [CrossRef]
  68. Gherasim, C.V.; Hancková, K.; Palarčík, J.; Mikulášek, P. Investigation of cobalt(II) retention from aqueous solutions by a polyamide nanofiltration membrane. J. Membr. Sci. 2015, 490, 46–56. [Google Scholar] [CrossRef]
  69. Murthy, Z.V.P.; Chaudhari, L.B. Application of nanofiltration for the rejection of nickel ions from aqueous solutions and estimation of membrane transport parameters. J. Hazard. Mater. 2008, 160, 70–77. [Google Scholar] [CrossRef] [PubMed]
  70. Boricha, A.G.; Murthy, Z.V.P. Preparation, characterization and performance of nanofiltration membranes for the treatment of electroplating industry effluent. Sep. Purif. Technol. 2009, 65, 282–289. [Google Scholar] [CrossRef]
  71. Gherasim, C.V.; Cuhorka, J.; Mikulášek, P. Analysis of lead(II) retention from single salt and binary aqueous solutions by a polyamide nanofiltration membrane: Experimental results and modelling. J. Membr. Sci. 2013, 436, 132–144. [Google Scholar] [CrossRef]
  72. Gao, J.; Sun, S.P.; Zhu, W.P.; Chung, T.S. Green modification of outer selective P84 nanofiltration (NF) hollow fiber membranes for cadmium removal. J. Membr. Sci. 2016, 499, 361–369. [Google Scholar] [CrossRef] [Green Version]
  73. Muthukrishnan, M.; Guha, B.K. Effect of pH on rejection of hexavalent chromium by nanofiltration. Desalination 2008, 219, 171–178. [Google Scholar] [CrossRef]
  74. Yu, Y.; Zhao, C.; Wang, Y.; Fan, W.; Luan, Z. Effects of ion concentration and natural organic matter on arsenic(V) removal by nanofiltration under different transmembrane pressures. J. Environ. Sci. 2013, 25, 302–307. [Google Scholar] [CrossRef]
  75. Urgun-Demirtas, M.; Benda, P.L.; Gillenwater, P.S.; Negri, M.C.; Xiong, H.; Snyder, S.W. Achieving very low mercury levels in refinery wastewater by membrane filtration. J. Hazard. Mater. 2012, 215–216, 98–107. [Google Scholar] [CrossRef]
  76. Figoli, A.; Cassano, A.; Criscuoli, A.; Mozumder, M.S.I.; Uddin, M.T.; Islam, M.A.; Drioli, E. Influence of operating parameters on the arsenic removal by nanofiltration. Water Res. 2010, 44, 97–104. [Google Scholar] [CrossRef] [PubMed]
  77. Murthy, Z.V.P.; Chaudhari, L.B. Separation of binary heavy metals from aqueous solutions by nanofiltration and characterization of the membrane using Spiegler-Kedem model. Chem. Eng. J. 2009, 150, 181–187. [Google Scholar] [CrossRef]
  78. Koseoglu, H.; Kitis, M. The recovery of silver from mining wastewaters using hybrid cyanidation and high-pressure membrane process. Miner. Eng. 2009, 22, 440–444. [Google Scholar] [CrossRef]
  79. Zeng, G.; He, Y.; Zhan, Y.; Zhang, L.; Pan, Y.; Zhang, C.; Yu, Z. Novel polyvinylidene fluoride nanofiltration membrane blended with functionalized halloysite nanotubes for dye and heavy metal ions removal. J. Hazard. Mater. 2016, 317, 60–72. [Google Scholar] [CrossRef]
  80. Zhu, W.P.; Gao, J.; Sun, S.P.; Zhang, S.; Chung, T.S. Poly(amidoamine) dendrimer (PAMAM) grafted on thin film composite (TFC) nanofiltration (NF) hollow fiber membranes for heavy metal removal. J. Membr. Sci. 2015, 487, 117–126. [Google Scholar] [CrossRef]
  81. Gao, J.; Sun, S.P.; Zhu, W.P.; Chung, T.S. Chelating polymer modified P84 nanofiltration (NF) hollow fiber membranes for high efficient heavy metal removal. Water Res. 2014, 63, 252–261. [Google Scholar] [CrossRef]
  82. Wang, K.Y.; Chung, T.S. Fabrication of polybenzimidazole (PBI) nanofiltration hollow fiber membranes for removal of chromate. J. Membr. Sci. 2006, 281, 307–315. [Google Scholar] [CrossRef]
  83. Ren, X.; Zhao, C.; Du, S.; Wang, T.; Luan, Z.; Wang, J.; Hou, D. Fabrication of asymmetric poly (m-phenylene isophthalamide) nanofiltration membrane for chromium(VI) removal. J. Environ. Sci. 2010, 22, 1335–1341. [Google Scholar] [CrossRef]
  84. Nedzarek, A.; Drost, A.; Harasimiuk, F.B.; Tórz, A. The influence of pH and BSA on the retention of selected heavy metals in the nanofiltration process using ceramic membrane. Desalination 2015, 369, 62–67. [Google Scholar] [CrossRef]
  85. Sayed, S.; Tarek, S.; Dijkstra, I.; Moerman, C. Optimum operation conditions of direct capillary nanofiltration for wastewater treatment. Desalination 2007, 214, 215–226. [Google Scholar] [CrossRef] [Green Version]
  86. Zolfaghari, G.; Kargar, M. Nanofiltration and microfiltration for the removal of chromium, total dissolved solids, and sulfate from water. MethodsX 2019, 6, 549–557. [Google Scholar] [CrossRef] [PubMed]
  87. Zhu, W.P.; Sun, S.P.; Gao, J.; Fu, F.J.; Chung, T.S. Dual-layer polybenzimidazole/polyethersulfone (PBI/PES) nanofiltration (NF) hollow fiber membranes for heavy metals removal from wastewater. J. Membr. Sci. 2014, 456, 117–127. [Google Scholar] [CrossRef]
  88. Lv, J.; Wang, K.Y.; Chung, T.S. Investigation of amphoteric polybenzimidazole (PBI) nanofiltration hollow fiber membrane for both cation and anions removal. J. Membr. Sci. 2008, 310, 557–566. [Google Scholar] [CrossRef]
  89. Gao, J.; Sun, S.P.; Zhu, W.P.; Chung, T.S. Polyethyleneimine (PEI) cross-linked P84 nanofiltration (NF) hollow fiber membranes for Pb2+ removal. J. Membr. Sci. 2014, 452, 300–310. [Google Scholar] [CrossRef]
  90. Qi, Y.; Zhu, L.; Shen, X.; Sotto, A.; Gao, C.; Shen, J. Polythyleneimine-modified original positive charged nanofiltration membrane: Removal of heavy metal ions and dyes. Sep. Purif. Technol. 2019, 222, 117–124. [Google Scholar] [CrossRef]
  91. Zareei, F.; Hosseini, S.M. A new type of polyethersulfone based composite nanofiltration membrane decorated by cobalt ferrite-copper oxide nanoparticles with enhanced performance and antifouling property. Sep. Purif. Technol. 2019, 226, 48–58. [Google Scholar] [CrossRef]
  92. Rafieian, F.; Jonoobi, M.; Yu, Q. A novel nanocomposite membrane containing modified cellulose nanocrystals for copper ion removal and dye adsorption from water. Cellulose 2019, 26, 3359–3373. [Google Scholar] [CrossRef]
  93. Gomes, S.; Cavaco, S.A.; Quina, M.J.; Gando-Ferreira, L.M. Nanofiltration process for separating Cr(III) from acid solutions: Experimental and modelling analysis. Desalination. 2010, 254, 80–89. [Google Scholar] [CrossRef]
  94. López, J.; Reig, M.; Gibert, O.; Cortina, J.L. Increasing sustainability on the metallurgical industry by integration of membrane nanofiltration processes: Acid recovery. Sep. Purif. Technol. 2019, 226, 267–277. [Google Scholar] [CrossRef]
  95. López, J.; Gibert, O.; Cortina, J.L. Evaluation of an extreme acid-resistant sulphonamide based nanofiltration membrane for the valorisation of copper acidic effluents. Chem. Eng. J. 2020, 127015. [Google Scholar] [CrossRef]
  96. Castro-Muñoz, R.; González-Melgoza, L.L.; García-Depraect, O. Ongoing progress on novel nanocomposite membranes for the separation of heavy metals from contaminated water. Chemosphere 2021, 270, 129421. [Google Scholar] [CrossRef]
  97. Shahalamqb, A.M.; Al-harthyb, A.; Al-zawhryb, A. Feed water pretreatment in RO systems in the Middle East. Desalination 2015, 37, 16–24. [Google Scholar]
  98. Rodrigues Pires da Silva, J.; Merçon, F.; Guimarães Costa, C.M.; Radoman Benjo, D. Application of reverse osmosis process associated with EDTA complexation for nickel and copper removal from wastewater. Desalin. Water Treat. 2016, 57, 19466–19474. [Google Scholar] [CrossRef]
  99. Petrinic, I.; Korenak, J.; Povodnik, D.; Hélix-Nielsen, C. A feasibility study of ultrafiltration/reverse osmosis (UF/RO)-based wastewater treatment and reuse in the metal finishing industry. J. Clean. Prod. 2015, 101, 292–300. [Google Scholar] [CrossRef]
  100. Mnif, A.; Bejaoui, I.; Mouelhi, M.; Hamrouni, B. Hexavalent Chromium Removal from Model Water and Car Shock Absorber Factory Effluent by Nanofiltration and Reverse Osmosis Membrane. Int. J. Anal. Chem. 2017, 2017. [Google Scholar] [CrossRef] [PubMed]
  101. Çimen, A. Removal of chromium from wastewater by reverse osmosis. Russ. J. Phys. Chem. A 2015, 89, 1238–1243. [Google Scholar] [CrossRef]
  102. Kurniawan, T.A.; Chan, G.Y.S.; Lo, W.H.; Babel, S. Physico-chemical treatment techniques for wastewater laden with heavy metals. Chem. Eng. J. 2006, 118, 83–98. [Google Scholar] [CrossRef]
  103. Zhong, C.-M.; Xu, Z.-L.; Fang, X.-H.; Cheng, L. Treatment of Acid Mine Drainage (AMD) by Ultra-Low-Pressure Reverse Osmosis and Nanofiltration. Environ. Eng. Sci. 2007, 24, 1297–1306. [Google Scholar] [CrossRef]
  104. Ricci, B.C.; Ferreira, C.D.; Aguiar, A.O.; Amaral, M.C.S. Integration of nanofiltration and reverse osmosis for metal separation and sulfuric acid recovery from gold mining effluent. Sep. Purif. Technol. 2015, 154, 11–21. [Google Scholar] [CrossRef]
  105. Choi, J.; Im, S.J.; Jang, A. Application of volume retarded osmosis—Low pressure membrane hybrid process for recovery of heavy metals in acid mine drainage. Chemosphere 2019, 232, 264–272. [Google Scholar] [CrossRef]
  106. Balanyà, T.; Labanda, J.; Llorens, J.; Sabaté, J. Influence of chemical speciation on the separation of metal ions from chelating agents by nanofiltration membranes. Sep. Sci. Technol. 2019, 54, 143–152. [Google Scholar] [CrossRef]
  107. Wei, X.Z.; Gan, Z.Q.; Shen, Y.J.; Qiu, Z.L.; Fang, L.F.; Zhu, B.K. Negatively-charged nanofiltration membrane and its hexavalent chromium removal performance. J. Colloid Interface Sci. 2019, 553, 475–483. [Google Scholar] [CrossRef]
  108. Sangeetha, K.; Sudha, P.N.; Faleh, A.A.; Sukumaran, A. Novel chitosan based thin sheet nanofiltration membrane for rejection of heavy metal chromium. Int. J. Biol. Macromol. 2019, 132, 939–953. [Google Scholar] [CrossRef]
  109. Jia, T.Z.; Lu, J.P.; Cheng, X.Y.; Xia, Q.C.; Cao, X.L.; Wang, Y.; Xing, W.; Sun, S.P. Surface enriched sulfonated polyarylene ether benzonitrile (SPEB) that enhances heavy metal removal from polyacrylonitrile (PAN) thin-film composite nanofiltration membranes. J. Membr. Sci. 2019, 580, 214–223. [Google Scholar] [CrossRef]
  110. Belkhouche, N.-E.; Merad, N.S.; Mesli, M.; Sefrou, Z. Separation of cobalt and nickel by nanofiltration using a FilmTec membrane. Euro-Mediterr. J. Environ. Integr. 2018, 3, 1–11. [Google Scholar] [CrossRef]
  111. Al-Rashdi, B.A.M.; Johnson, D.J.; Hilal, N. Removal of heavy metal ions by nanofiltration. Desalination 2013, 315, 2–17. [Google Scholar] [CrossRef]
  112. Maher, A.; Sadeghi, M.; Moheb, A. Heavy metal elimination from drinking water using nanofiltration membrane technology and process optimization using response surface methodology. Desalination 2014, 352, 166–173. [Google Scholar] [CrossRef]
  113. Abdi, G.; Alizadeh, A.; Zinadini, S.; Moradi, G. Removal of dye and heavy metal ion using a novel synthetic polyethersulfone nanofiltration membrane modified by magnetic graphene oxide/metformin hybrid. J. Membr. Sci. 2018, 552, 326–335. [Google Scholar] [CrossRef]
  114. Gherasim, C.V.; Mikulášek, P. Influence of operating variables on the removal of heavy metal ions from aqueous solutions by nanofiltration. Desalination 2014, 343, 67–74. [Google Scholar] [CrossRef]
  115. Yoon, J.; Amy, G.; Chung, J.; Sohn, J.; Yoon, Y. Removal of toxic ions (chromate, arsenate, and perchlorate) using reverse osmosis, nanofiltration, and ultrafiltration membranes. Chemosphere 2009, 77, 228–235. [Google Scholar] [CrossRef]
  116. Harisha, R.S.; Hosamani, K.M.; Keri, R.S.; Nataraj, S.K.; Aminabhavi, T.M. Arsenic removal from drinking water using thin film composite nanofiltration membrane. Desalination 2010, 252, 75–80. [Google Scholar] [CrossRef]
  117. Sen, M.; Manna, A.; Pal, P. Removal of arsenic from contaminated groundwater by membrane-integrated hybrid treatment system. J. Membr. Sci. 2010, 354, 108–113. [Google Scholar] [CrossRef]
  118. Wei, X.; Kong, X.; Wang, S.; Xiang, H.; Wang, J.; Chen, J. Removal of heavy metals from electroplating wastewater by thin-film composite nanofiltration hollow-fiber membranes. Ind. Eng. Chem. Res. 2013, 52, 17583–17590. [Google Scholar] [CrossRef]
  119. Ku, Y.; Chen, S.W.; Wang, W.Y. Effect of solution composition on the removal of copper ions by nanofiltration. Sep. Purif. Technol. 2005, 43, 135–142. [Google Scholar] [CrossRef]
  120. Mehdipour, S.; Vatanpour, V.; Kariminia, H.R. Influence of ion interaction on lead removal by a polyamide nanofiltration membrane. Desalination 2015, 362, 84–92. [Google Scholar] [CrossRef]
  121. Chaudhari, L.B.; Murthy, Z.V.P. Separation of Cd and Ni from multicomponent aqueous solutions by nanofiltration and characterization of membrane using IT model. J. Hazard. Mater. 2010, 180, 309–315. [Google Scholar] [CrossRef] [PubMed]
  122. Tanninen, J.; Platt, S.; Weis, A.; Nyström, M. Long-term acid resistance and selectivity of NF membranes in very acidic conditions. J. Membr. Sci. 2004, 240, 11–18. [Google Scholar] [CrossRef]
  123. Chaabane, T.; Taha, S.; Taleb Ahmed, M.; Maachi, R.; Dorange, G. Removal of copper from industrial effluent using a spiral wound module—Film theory and hydrodynamic approach. Desalination 2006, 200, 403–405. [Google Scholar] [CrossRef]
  124. Sudilovskiy, P.S.; Kagramanov, G.G.; Kolesnikov, V.A. Use of RO and NF for treatment of copper containing wastewaters in combination with flotation. Desalination 2008, 221, 192–201. [Google Scholar] [CrossRef]
  125. Nguyen, V.T.; Vigneswaran, S.; Ngo, H.H.; Shon, H.K.; Kandasamy, J. Arsenic removal by a membrane hybrid filtration system. Desalination 2009, 236, 363–369. [Google Scholar] [CrossRef] [Green Version]
  126. Mohsen-Nia, M.; Montazeri, P.; Modarress, H. Removal of Cu2+ and Ni2+ from wastewater with a chelating agent and reverse osmosis processes. Desalination 2007, 217, 276–281. [Google Scholar] [CrossRef]
  127. ZHANG, L.; WU, Y.; QU, X.; LI, Z.; NI, J. Mechanism of combination membrane and electro-winning process on treatment and remediation of Cu2+ polluted water body. J. Environ. Sci. 2009, 21, 764–769. [Google Scholar] [CrossRef]
  128. Abejón, A.; Garea, A.; Irabien, A. Arsenic removal from drinking water by reverse osmosis: Minimization of costs and energy consumption. Sep. Purif. Technol. 2015, 144, 46–53. [Google Scholar] [CrossRef]
  129. Teychene, B.; Collet, G.; Gallard, H.; Croue, J.P. A comparative study of boron and arsenic (III) rejection from brackish water by reverse osmosis membranes. Desalination 2013, 310, 109–114. [Google Scholar] [CrossRef]
  130. Akin, I.; Arslan, G.; Tor, A.; Cengeloglu, Y.; Ersoz, M. Removal of arsenate [As(V)] and arsenite [As(III)] from water by SWHR and BW-30 reverse osmosis. Desalination 2011, 281, 88–92. [Google Scholar] [CrossRef]
  131. Qdais, H.A.; Moussa, H. Removal of heavy metals from wastewater by membrane processes: A comparative study. Desalination 2004, 164, 105–110. [Google Scholar] [CrossRef]
  132. Shirazi, S.; Lin, C.J.; Chen, D. Inorganic fouling of pressure-driven membrane processes—A critical review. Desalination 2010, 250, 236–248. [Google Scholar] [CrossRef]
  133. Tu, K.L.; Chivas, A.R.; Nghiem, L.D. Effects of membrane fouling and scaling on boron rejection by nano fi ltration and reverse osmosis membranes. Desalination 2011, 279, 269–277. [Google Scholar] [CrossRef] [Green Version]
  134. Jafari, M.; Vanoppen, M.; van Agtmaal, J.M.C.; Cornelissen, E.R.; Vrouwenvelder, J.S.; Verliefde, A.; van Loosdrecht, M.C.M.; Picioreanu, C. Cost of fouling in full-scale reverse osmosis and nanofiltration installations in the Netherlands. Desalination. 2021, 500, 114865. [Google Scholar] [CrossRef]
  135. Lyu, Z.; Ng, T.C.A.; Duc, T.T.; Lim, G.J.H.; Gu, Q.; Zhang, L.; Zhang, Z.; Ding, J.; Thien, N.P.; Wang, J.; et al. 3D-printed surface-patterned ceramic membrane with enhanced performance in crossflow filtration. J. Membr. Sci. 2020, 606, 118138. [Google Scholar] [CrossRef]
  136. Ng, T.C.A.; Lyu, Z.; Gu, Q.; Zhang, L.; Poh, W.J.; Zhang, Z.; Wang, J.; Ng, H.W. Effect of gradient profile in ceramic membranes on filtration characteristics: Implications for membrane development. J. Membr. Sci. 2020, 595, 117576. [Google Scholar] [CrossRef]
  137. Zhang, L.; Ng, T.C.A.; Liu, X.; Gu, Q.; Pang, Y.; Zhang, Z.; Lyu, Z.; He, Z.; Ng, H.Y.; Wang, J. Hydrogenated TiO2 membrane with photocatalytically enhanced anti-fouling for ultrafiltration of surface water. Appl. Catal. 2020, 264, 118528. [Google Scholar] [CrossRef]
Figure 1. Various pressure-driven membrane processes.
Figure 1. Various pressure-driven membrane processes.
Processes 09 00752 g001
Figure 2. Process of Micellar Enhanced Filtration.
Figure 2. Process of Micellar Enhanced Filtration.
Processes 09 00752 g002
Table 1. Various membranes in wastewater remediation.
Table 1. Various membranes in wastewater remediation.
Membrane MaterialCharacteristic of MembraneHeavy Metal TargetedSurfactant/Complexing Agent UsedOptimum Pressure (Bar)Surfactant Concentration (mM)Initial Concentration (mg/L)pH% RemovalReference
CeramicMWCO = 210 kDaNi(II)
Co(II)
Sodium dodecyl sulphate2.8 0.02510 753
51
[20]
PAN MembraneArea = 0.00124 m2As(V)Cetyl Pyridinium Chloride (CPC) 1 5 17–896.9[21]
Polyether sulphoneMWCO = 6000 g mol−1
Area = 0.3 m2
TMP <= 0.15 MPa
Cd(II)
Zn(II)
Sodium dodecyl sulphate0.72.1550 92–98[22]
Amicon regenerated celluloseMWCO = 10 kDaCd(II)
Zn(II)
Sodium dodecyl sulphate313.9
14.2
23 99[23]
PolycarbonateTMP = 250 kPaNi(II)Sodium lauryl ether sulphate 2 9.2 98.6[24]
Polyether sulphoneMWCO = 10 kDa
Area = 9.6 cm2
Permeate flux = 150 L.m2/h at 0.35 MPa
Cu(II)
Cd(II)
Zn(II)
Pb(II)
Sodium dodecyl sulphate 8(50–300)>3
>3
3–10
3–10
99[25]
Polyether sulphoneMCO = 10 kDa
Area = 32.15 × 10−4 m2
Cd(II)Rhamnolipid2.768.04607.892[26]
CeramicMWCO = 1 kDaZn(II)Sodium dodecyl sulphate0.810 2 99[27]
Polyacrylonitrile (PAN)MWCO = 300,000
Area = 4.8 m2
Zn(II)Sodium dodecyl sulphate20.21 19.32784.67[28]
Polyether sulphoneMWCO = 10 kDa
Area = 1.6 m2
Cd(II)
Cu(II)
Sodium dodecyl sulphate3 60 0.37
0.41
85
81
[29]
PolysulphoneMWCO =10 k Da
Area = 0.014 m2
Cr(VI)
Cr(III)
Rhamnolipid0.70.0210 698.7
96.2
[30]
Regenerated CelluloseMWCO = 10 kDa
Are a= 0.0013 m2
Cu(II)
Cd(II)
Zn(II)
Ni(II)
Mg(II)
nonaoxyethylene oleylether carboxylic acid (RO90)330.439206.5>95%[31]
Polysulphone MWCO = 1 kDa
Area = 0.004 m2
Ni(II)Sodium dodecyl sulphate2.51610797%[32]
Polyacrylonitrile (PAN)MWCO = 100 kDa
Area = 0.07 m2
Ni(II)
Zn(II)
Sodium dodecyl sulphate112.75 23796.3
96.7
[33]
PolysulphoneMWCO = 10 kDa
Area = 0.004 m2
Ni(II)Sodium dodecyl sulphate18101199[34]
PolysulphoneMWCO = 10 kDa
Area = 0.3 m2
Cd(II)Sodium dodecyl sulphate0.380.45 97[35]
Polyether sulphoneMWCO = 5 kDa, 10 kDa, 30 kDa
Area = 0.00096 m2
Cd(II)Sodium dodecyl sulphate1 4 10 90[36]
Polyether sulphoneMWCO = 8 kDa
Area = −0.005 m2
Cd(II)Sodium dodecyl sulphate 7.3350 98.4[37]
HydrophilicMWCO = 10 kDaCu(II)polyoxyethylene Octyl phenyl ether (Triton-X) plus Sodium dodecyl sulphate 2.0961.29
5.67
9.2592[38]
Polyether sulfoneMWCO = 10 kDa
Area = 0.003019 m2
Cd(II)
Cu(II)
Pb(II)
Zn(II)
Sodium dodecyl sulphate19 10 >90[16]
Table 2. Effect of environmental chemistry on the removal of heavy metals by membranes.
Table 2. Effect of environmental chemistry on the removal of heavy metals by membranes.
MembraneCharacteristic of Membrane (MWCO)Heavy Metal Surfactant/Complexing Agent UsedOptimum Pressure (bar) Surfactant ConcentrationInitial Concentration (mg/L)pH% RemovalRef
Ceramic15 kDaCu(II)Poly (acrylic acid) sodium3 1 wt%1604–599.5[48]
Ceramic15,000 g/molCr(III)Polyvinyl alcohol (PVA) 1 wt%925>90%[49]
Polyether sulphone10 kDaPb(II)
Cu(II)
Fe(III)
Polyvinylamine20.1 wt%25 799
97
99
[40]
Ceramic10 kDaCu(II)Poly (acrylic acid) 0.4 wt%160 5.599.5[50]
Ceramic10 kDaCu(II)
Zn(II)
Partially ethoxylated polyethyleneimine (PEPEI)3 0.06 wt%90.626Selectivity ratio Cu(II)/Zn(II) = 12.31[51]
Polyether sulphone10 kDaCd(II)Poly (ammonium acrylate)2 3.71 × 10−4 mol/L466.3299[52]
Cellulose10 kDaCu(II)
Zn(II)
Poly (acrylic acid)3 2 × 10−3 mol/L46597
75
[53]
Thin Film Composite3.5 kDaNi(II)Chitosan28 2 × 10−2 mol/L0.0725.490[54]
Polyether sulphone10 kDaHg(II)Polyvinylamine20.05 wt%10 >90[55]
Polyether sulphone60 kDaCu(II)Polyethylenimine (PEI)1.7 25 mM230394[56]
Polysulphone8 kDa, 15 kDaCd(II)Poly(ammonium) acrylate2 46498[57]
Ceramic10 kDaCr(VI)poly(diallyldimethylammonium chloride) (PDADMAC)4 0.1 wt%50 999[58]
Polyethersulphone10 kDaCu(II)
Ni(II)
Cr(III)
Carboxy methyl cellulose1 1 g/L10 797.6
99.1
99.5
[59]
Polyether sulphone10 kDaHg(II)Polyvinylamine20.1 wt%10 6–799[60]
Ceramic10 kDaCu(II)Partially ethoxylated polyethylenimine (PEPEI)4 0.06 wt%208 mg Cu/g PEPEI697[61]
Ceramic10 kDaPb(II)Poly(acrylic) acid (PAA)4 0.036%100 6100[62]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Verma, B.; Balomajumder, C.; Sabapathy, M.; Gumfekar, S.P. Pressure-Driven Membrane Process: A Review of Advanced Technique for Heavy Metals Remediation. Processes 2021, 9, 752. https://doi.org/10.3390/pr9050752

AMA Style

Verma B, Balomajumder C, Sabapathy M, Gumfekar SP. Pressure-Driven Membrane Process: A Review of Advanced Technique for Heavy Metals Remediation. Processes. 2021; 9(5):752. https://doi.org/10.3390/pr9050752

Chicago/Turabian Style

Verma, Bharti, Chandrajit Balomajumder, Manigandan Sabapathy, and Sarang P. Gumfekar. 2021. "Pressure-Driven Membrane Process: A Review of Advanced Technique for Heavy Metals Remediation" Processes 9, no. 5: 752. https://doi.org/10.3390/pr9050752

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop