Article pubs.acs.org/IECR
Inorganic Membranes for the Recovery of Effluent from Municipal Wastewater Treatment Plants Ali Farsi, Sofie Hammer Jensen, Peter Roslev, Vittorio Boffa, and Morten Lykkegaard Christensen* Department of Chemistry and Bioscience, Aalborg University, Fredrik Bajers Vej 7H, DK-9220 Aalborg East, Denmark ABSTRACT: Effluent from the secondary treatment of municipal wastewater was treated with membrane filtration to reduce its toxicity. Five different inorganic membranes were tested, namely, an α-alumina microfiltration membrane, an anatase titania ultrafiltration membrane, a γ-alumina nanofiltration membrane, an amorphous titania nanofiltration membrane, and an amorphous hybrid organo-silica membrane. The permeabilities and selectivities (color, UV254-absorbing components, conductivity and inorganic nitrogen compounds) of the membranes were determined, and the γ-alumina nanofiltration membrane was found to be the most promising membrane for the treatment of the effluent. The effluent flux was measured to be approximately 40 L m−2 h−1 for the γ-alumina nanofiltration membrane, and it removed nearly 75% of the UV254-absorbing components and 15% of the ions. It also removed 40% of the CuCl and 25% of the CuSO4 from the spiked effluent. The fouling resistance was less pronounced for the γ-alumina membrane compared with the other membranes. The removal of fecal indicator bacteria was determined by measuring the amounts of Escherichia coli and Enterococci, and the removal of toxic compounds was investigated in bioassays with Daphnia magna and Aliivibrio fischeri. The γ-alumina nanofiltration membrane reduced the wastewater concentration of E. coli (97.3%) and Enterococci (98.4%), and the bioassays demonstrated that filtration with the γalumina nanofiltration membrane reduced the overall toxicity of the effluent.
1. INTRODUCTION Energy-efficient technologies are essential for water cleaning because water shortages are becoming increasingly common. Existing wastewater treatment plants (WWTPs) remove organic materials and nutrients (i.e., nitrogen and phosphorus) from the wastewater. However, WWTP effluents often contain toxic compounds, such as organic micropollutants (OMPs), dissolved inorganic nitrogen compounds (DINs), and heavy metal ions.1−4 Such OMPs can accumulate in aquatic organisms and adversely affect their growth and reproduction even though their concentrations are typically in the milligram/nanogram per liter range or lower.4,5 Based on their functions, the OMPs have been composed of pharmaceuticals and personal care products, steroid hormones, endocrine-disrupting compounds, surfactants, flame retardants, pesticides, synthetic fragrances, industrial additives, and many other emerging compounds.3−6 The amounts of these contaminants that have been found in the environment have been increasing, and their extremely low concentrations, as well as their bio-persistence and bioaccumulation, have rendered their measurement and subsequent treatment difficult. Luo et al.3 provided a comprehensive review of the occurrence data of OMPs in WWTP effluent from recent studies. DINs exist in fairly high concentrations (typically > 5 mg N/ L) and in various chemical forms, including ammonium (NH4+), nitrate (NO3−), and nitrite (NO2−).7,8 The DINs in WWTP effluent can stimulate bacterial and phytoplankton growth in the receiving waters.9 The DINs can also associate to increase the concentration of hydrogen ions in freshwater ecosystems, resulting in the acidification of the systems and thus inducing the occurrence of toxic algae, which can reach toxic levels that impair the abilities of the aquatic animals to survive, grow and reproduce. Ingested nitrites and nitrates can also be harmful for human health.10,11 Most of the heavy metals © XXXX American Chemical Society
in wastewater, such as copper, nickel, chromium, zinc, and silver, are resistant to biodegradation, and they have a propensity for bioaccumulation in living organisms, causing serious health problems.12,13 Ingested copper may cause nausea, vomiting, and liver and kidney damage.14 Therefore, WWTP effluent must be treated before discharge. Certain advanced physical, chemical, and biological technologies and methods have been investigated to assess their effectiveness for reducing the toxicity of WWTP effluents. These technologies and methods include coagulation−flocculation,15 advanced oxidation processes,16−18 precipitation,19 sorption,20 membrane bioreactor,21 ion exchange,22 sand filtration, and activated carbon adsorption.23 Among these methods, membrane processes, and particularly pressure-driven membrane processes, are promising because no heating or chemical additives are required.24−27 Previous studies28,29 have shown that the costs involved in the operation of the membrane systems in the recovery of WWTP effluents were competitive with conventional treatment processes, such as chlorine and ozone processes, but until recently, the vast majority of membranes that have been tested were polymeric membranes.29−31 One alternative to the use of polymeric membranes is inorganic membranes. Inorganic membranes have been used in many applications, such as milk microfiltration, beer clarification, food processing, and in the oil and petrochemical industry.32−34 Compared to polymeric membranes, inorganic membranes are more expensive and have a lower loading Received: January 6, 2015 Revised: March 9, 2015 Accepted: March 18, 2015
A
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Article
Industrial & Engineering Chemistry Research density; however, they have demonstrated high chemical stability, large permeability, and low tendencies for fouling.35,36 Therefore, inorganic membranes may be useful in the treatment of municipal WWTP effluent. In this study, five inorganic membranes with different pore sizes and different materials were tested. These membranes covered a broad range of applications, ranging from microfiltration (MF) to nanofiltration (NF). MF and ultrafiltration (UF) membranes were expected to have pore sizes that were not sufficiently small for the retention of OMPs and that would render them unsuitable for this application; however, these membranes have been used as supports for the NF membranes and were thus included in this study. The membrane permeabilities and the retention of the colors, UV254-absorbing components (UVAs), conductivities, and DINs were determined for all membranes. There was a correlation between the abatement of various OMPs and the corresponding losses in the UVAs,6,37−39 as well as between the removal of the total ions and the decrease in the conductivities. The color measurements that were obtained through a spectrophotometric method were a useful index of the dissolved humic-like substances in water.6,37 Additionally, Wert et al.6 proposed the reduction of color as a potential method to assess the removal of pharmaceuticals. The membranes that provided more than 30 L m−2 h−1 (LMH) of uncolored permeate flux and a 75% retention of UVAs were selected for further investigation. The removal of the toxic compounds and the indicator bacteria by the optimum membrane were investigated using two bioassays that targeted the inhibition of Daphnia magna and Aliivibrio fischeri and that quantified the indicator bacteria Escherichia coli and Enterococci.
Table 1. Structure of the Five Membranes Used This Study no.
name
1
MF αalumina UF titania NF γalumina NF titania Hybsi
2 3 4 5
active layer
interlayer
support
α-alumina
microfiltration
anatase titania γ-alumina amorphous titania amorphous organo-silica
type of application
UF titania NF γalumina
MF αalumina MF αalumina MF αalumina MF αalumina
ultrafiltration nanofiltration nanofiltration pervaporation
2.3. Experimental Setup. The experimental cross-flow filtration setup is shown in Figure 1. A feed solution was pumped into the membrane by a feed pump (BEVI, IEC 34-1, Sweden) that was capable of providing pressures of up to 1.9 MPa. The mass flow of the permeate was measured by a balance (Mettler Toledo, Mono Bloc series, Switzerland) connected to a computer. The feed pressure was measured before and after the membrane by two pressure transmitters (Danfoss, MBS 4010, Denmark), and an electronic heat sensor (Kamstrup A/S, Denmark) was used measure the feed temperature before the membrane module. The cross-flow stream was provided by a rotary lobe pump (Philipp Hilge Gmbh & Co, Novalobe 60/1.90, Germany) that was capable of generating a cross-flow of 2 L/min. The cross-flow rate was measured by a microprocessor-based flow rate transmitter (Siemens, MAG 50000). The retentate stream was controlled by a manual valve (Nupro ). 2.4. Filtration Protocol. Prior to each experiment, the membrane was tested for 2 h at transmembrane pressure (TMP) = 6 bar using deionized water to measure the clean water flux. Then, the WWTP samples were filtered at room temperature and the effluent flux through the membrane was measured. TMP for all experiments and membranes was 6 bar. The recovery rate (ratio of permeate flow to feed flow) was approximately 50%. The obtained permeate was characterized immediately using a pH meter (Radiometer PHM 92 Lab pHmeter) and viscometer (Brookfield, Model DV-II+), and then, the samples were kept at −4 °C and −18 °C for the analytical measurements and bioassays, respectively. The cross-flow rate was maintained at 0.75 L/min to provide a cross-flow velocity of approximately 20 m/s. The cross-flow was high compared with the normal operations to reduce the effect of the concentration polarization.35 2.5. Membrane Permeability. The membrane permeability (Lp) was determined from eq 1:
2. MATERIALS AND METHODS 2.1. Sampling. Samples of the effluent from the secondary wastewater treatment were collected from a municipal WWTP (250 000 PE, Aalborg West, Denmark) in sterilized 4 L glass bottles. All samples were immediately filtered by glass fiber filters (0.45 μm) to eliminate the suspended solids and subsequently stored at 4 °C to minimize changes in the constituents in the water. The Aalborg West WWTP effluent contained approximately 2.8 mg/L organic matters, 5 mg/L inorganic nitrogen compounds and it conductivity was measured to be 1120 μS/cm.40 Because sodium and chloride are the main inorganic ions in water, we assumed it as a dilute NaCl solution. 2.2. Membranes. Five different monotubular membranes (250 × 10 × 7 mm (L × OD × ID)) were purchased from Pervatech B.V. (The Netherlands) and directly used for filtration. All five of the membranes possessed an asymmetric structure that consisted of an active layer and support layer. Table 1 lists the support and interlayer structures, the active layer compositions and the commercial uses for each membrane. For clarity, membranes will now be referenced according to their designation in Table 1. Furthermore, the nominal pore sizes were determined using the molecular weight cutoff method for all of the active layers, except Hybsi. The single gas permeation method35,41 was used to determine the pore size distribution of the Hybsi active layer, which was a challenge because some of their pores were not accessible to most of the gas probes. The Hybsi active layer was comprised of a network of hybrid silica chains that created an apparent pore size. Previous literature has suggested that the Hybsi membrane pores possessed a mean radius (rp) of between 0.15 and 0.3 nm.35,41,42
Lp =
Jp ΔPeff
(1)
where Jp was the measured permeate flux and ΔPeff was the effective pressure driving force (ΔP − Δπ). For dilute solutions, the osmotic pressure (Δπ) was calculated using the Van’t Hoff equation.43 As mentioned previously, the sample was considered to be a dilute sodium chloride solution (1120 μS/cm); the concentration of sodium chloride (c) was estimated from the NaCl molar conductivity (ΛNaCl), and the electrical conductivity (σ) of the sample was determined by eq 2 and was approximately 11 mM:44 B
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Article
Industrial & Engineering Chemistry Research
Figure 1. Experimental cross-flow filtration setup (TR, PR, and FR are temperature, pressure, and flow rate transmitters, respectively).
c=
σ ΛNaCl
(2)
The permeate flux was modeled using the modified Hagen− Poiseuille (HP) equation as follows: rp2ε ΔPeff Jp = 8ηappτ Δx
(3)
where rp, Δx, ε, and τ are the pore radius, membrane thickness, porosity, and tortuosity, respectively. The membrane materials were charged, and charge effects may have influenced the permeate flux due to the small pores of the NF membranes. To account for this electroviscous effect, an apparent viscosity (ηapp) was used instead of the bulk viscosity. The apparent viscosity was calculated using eq 4:35,45 ηapp ηb
⎡ ⎛ ⎢ 8β ⎜1 − ⎝ = ⎢1 − ⎢ ⎛ ⎛ (κrp)2 ⎜1 − β ⎜1 − ⎢ ⎝ ⎝ ⎣
Figure 2. Schematic representation of solvent transport through the concentration polarization and membrane layers.
⎤−1 ⎥ ⎥ I12(κrp) ⎞⎞ ⎥ − 2 ⎟⎟ ⎥ I0 (κrp) ⎠⎠ ⎦
⎞ κrpI0(κrp) ⎠ 2I1(κrp)
2I1(κrp) κrpI0(κrp)
thickness of the CP layer (δ) was calculated from eq 6 by assuming that the system was in a steady state.
⎟
δ= 7:
where ηb is the bulk viscosity of the solvent, I0 and I1 are the zero-order and first-order modified Bessel functions of the first kind, and κrp is a dimensionless number that indicates the ratio between the pore radius and electrical double layer thickness (κ−1). Furthermore, the dimensionless parameter β was determined by
Jp ⎛ exp k ⎜ cm =⎜ cb ⎜ R + (1 − R )exp ⎝
()
Jp
() k
⎞ ⎟ ⎟ ⎟ ⎠
(7)
where D is the diffusion coefficient of the solute, k is the mass transfer coefficient, cb is the bulk concentration, cm is the concentration on the membrane surface, and R is the solute retention. Moreover, the mass transfer coefficient, which is dependent on the flow regime, was calculated for the tube geometry under the turbulent regime (Re > 4000) as follows D D k = Sh = (0.04Re 0.75Sc 0.33) dh dh (8)
(ϵpϵ0E0κ )2 16π 2ηbσp
(6)
The concentration modulus (cm/cb) was calculated using eq
(4)
β=
D k
(5)
where ϵp is the average dielectric constant in the pore that was calculated based on the dielectric constant of the bulk solution as reported elsewhere,35 E0 is the surface potential, and σp is the electrical conductivity of the permeate. The surface potential (E0) was assumed to be equal to the ζ-potential.45 2.6. Concentration Polarization. Some of the solutes were retained by the membrane; these solutes accumulated and formed a highly concentrated layer near the membrane surface. The concentration polarization (CP) increased the osmotic pressure at the membrane surface and thereby lowered the permeate flux according to eq 3. Figure 2 represents a schematic of the concentration profile in the CP layer. The
where Sh, Re, and Sc are the Sherwood, Reynolds, and Schmidt numbers, respectively, and dh is the hydraulic diameter. 2.7. Active Layer Resistance. The total resistance of the membrane was calculated from the permeate flux: Lp R0 = η (9) C
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Article
Industrial & Engineering Chemistry Research Table 2. Model Parameters Used in This Study for Different Active Layers membrane parameter nominal pore size (dp) [nm] active layer thickness (l) [μm] active layer porosity (ε) [−] active layer tortuosity (τ) [−] |ζ|c [mV] resistance [m−1] Ro Rsu Rin Rac a
MF α-alumina
UF titania
NF γ-alumina
NF titania
Hybsi
110a 100 (100−300)b 0.36 (0.3−0.43)b 1.7 (1.5−2.5)b 35
15a 3 (0.4−5)b 0.38 (0.35−0.4)b 3 (2−3)b 15
4.4a 1 (1−2)b 0.55 (0.4−0.55)b 2.6 (2.5−15)b 60
2a 0.1 (0.1−0.4)b 0.42 (0.3−0.4)b 2.5 (2.5−5)b 15
0.4a 0.2 (0.1−0.5)b 0.2 (0.2−0.3) 11 (−) 20
1/Lp,1ηapp
1/Lp,2ηapp R1
1/Lp,3ηapp R1
R1
R2 = Ro − R1
R3 = Ro − R1
1/Lp,4ηapp R1 R2 Ro − R2 − R1
1/Lp,5ηapp R1 R3 Ro − R3 − R1
Data provided by manufacture. bData provided by literature.35,41,42,46−54 cAt pH 5.5 for a dilute solution (c < 0.001 M).
magna were determined according to ISO 634155 and ISO 10706.56 Female D. magna were obtained from a laboratory clone originating from ephippia (MicroBioTests Inc., Belgium). Prior to the tests, D. magna were cultivated at 20 °C for >10 generations in synthetic freshwater in 20 L tanks with a 16:8 h light/dark cycle. In the acute test (ISO 6341), 20 juvenile D. magna ( 436 nm). They reported that the membrane permeability of this NF titania membrane in the presence of wastewater varied between 5 and 13 LMH·bar−1, in T > 50 °C and ucf > 4.5 m/s, whereas Marchetti et al.66 showed that the clean water permeability of NF titania membrane (dp ∼ 0.9 nm, HITK/Inopore, Hermsdorf, Germany) is approximately 20 LMH·bar−1, at T = 20 °C. Lin et al.67 have studied the removal of pharmaceuticals and personal care products (PPCP) by cross-flow filtration using a polymeric NF membrane (NF90, dp ∼ 0.68 nm, manufactured by Dow FilmTech) at TMP = 6.8 bar. At pH = 8, i.e., our WWTP sample pH, clean water permeability was 10.6 LMH· bar−1, the flux declines were 53%, and rejection of organic PPCPs (molecular weight of 206−289 g/mol) was approximately 55−80%. Studied NF γ-alumina membrane in this work showed a better performance in this pH, i.e., clean water permeability = 12.8 LMH·bar−1, flux decline = 45% and 75% rejection of UVA254 component. The rejection of a specific toxic ion, copper, was tested by the NF γ-alumina membrane. The NF γ-alumina membrane rejected approximately 40% of the CuCl (Cu(I)) and 25% of the CuSO4 (Cu(II)) during the cross filtration of the spiked WWTP effluent whose copper concentration was set at approximately 1 ± 0.1 mg L−1. These results were consistent with our previous work, which demonstrated that the NF γalumina membrane removed approximately 20% of the NaCl and 40% of the MgCl2 from the dilute electrolyte at the same operational conditions.35 3.5. Bacteria and Toxicity Removal. The NF γ-alumina membrane dramatically reduced the concentrations of E. coli and the intestinal Enterococci in the permeate (Figure 8). The
Figure 9. Acute toxicity to D. magna of WWTP effluent and NF γalumina permeate (a), and acute toxicity to A. fischeri of CuCl spiked WWTP effluent and NF γ-alumina permeate (b).
and A. fischeri organisms suggested EC50 values above 50% (v/ v), which indicated a low acute toxicity of the NF γ-alumina membrane permeate (Figure 9). The NF γ-alumina membrane treatment was also capable of reducing the toxicity of the WWTP effluent with elevated concentrations of toxic ions. In this case, CuCl was added to the WWTP effluent prior to the NF γ-alumina membrane treatment to increase the background toxicity. The reduction of the toxicity due to the NF γ-alumina membrane treatment increased with the increasing wastewater concentrations, and a maximum reduction of 47−58% was obtained. This attenuation of toxicity was in the same range as the 40% reduction in the CuCl concentration observed before. The long-term incubation of A. fischeri for 3−7 h resulted in a detectable inhibition, but the effect was lower than what was observed for the nonrecovered WWTP effluent (Figure 10). The WWTP effluent inhibited A. fischeri by 28−38%, whereas the NF γ-alumina membrane reduced the inhibition of A. fischeri by 20−35% (Figure 10). These results supported the
Figure 8. Concentration of E. coli and enterococci in the WWTP effluent and NF γ-alumina permeate.
reduction in the MPN concentration was comparable with the removal efficiencies of 97.3% and 98.5% for E. coli and Enterococci, respectively. The removal of the fecal indicator bacteria from the WWTP effluent was relevant because these bacteria are now the guiding parameters in the testing of recreational water quality in many countries.68 Therefore, the filtration of the WWTP effluent may have contributed to the H
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research
Article
■
ACKNOWLEDGMENTS This research was carried out under project 0-59-11-1 from Danish National Advanced Technology Foundation, which is kindly acknowledged for financial support.
■
Figure 10. Inhibition of A. fischeri after 3, 5, and 7 h of exposure to WWTP effluent and NF γ-alumina permeate.
observations that the toxicity of the effluent was reduced during the NF γ-alumina membrane treatment.
4. CONCLUSIONS Different inorganic membranes, namely, a MF α-alumina membrane, UF titania membrane, NF γ-alumina membrane, NF titania membrane, and Hybsi membrane, were studied to test their ability to remove toxic compounds, including aromatic components, humic-like substances, OMPs, DINs and heavy metal ions, from WWTP effluent. The permeabilities and selectivities of the membranes were determined. The NF γalumina membrane was the most promising membrane for the recovery of WWTP effluent with regard to its permeate flux and selectivity. The membrane permeability was measured to be approximately 12.8 and 6.6 LMH·bar−1 for the NF γ-alumina membrane in the presence of deionized water and WWTP effluent, respectively. The NF γ-alumina membrane removed nearly 75% of the UVAs and 15% of the ions. The membrane rejected 40% of the CuCl and 25% of the CuSO4 from the spiked WWTP effluent. The overall resistance of the NF γalumina membrane active layer in the presence of the WWTP effluent was 33 × 1012 m−1, which was formed by a fouling resistance of nearly 20% and an active layer resistance of 80% by itself. The removal of the indicator bacteria and toxic compounds by the NF γ-alumina membrane were tested using bioassays that targeted E. coli, Enterococci, D. magna, and A. fischeri. Results from the bioassays indicated that the treatment with the NF γ-alumina membrane reduced the overall bacterial load and environmental toxicity of the treated water. Due to the permeability, selectivity and fouling performance of the NF γalumina membrane, this membrane should be considered as a promising alternative in the removal of both toxic organics and OMPs from the effluent of WWTPs.
■
■
NOMENCLATURE c = Concentration [mol m−3] cb = Bulk concentration [mol m−3] cm = Concentration on the membrane surface [mol m−3] D = Bulk diffusion coefficient [m2 s−1] dh = Hydraulic diameter [m] dp = Membrane nominal pore size [nm] E0 = Surface potential [V] I = Inhabitation in eq 11 [−] I0 = Zero-order modified Bessel function of the first kind [−] I1 = First order modified Bessel function of the first kind [−] Jp = Solvent flux [L m−2 h−1] k = Mass transfer coefficient [m s−1] Lp = Overall permeability [L m−2 h−1 bar−1] l = Active layer thickness [μm] R = Solute retention [-] Rac = Active layer resistance [m−1] Rac,e = Active layer resistance for WWTP effluent [m−1] Rac,w = Active layer resistance for deionized water [m−1] Ri = Measurement for inhabited sample in eq 11 [−] Rc = Measurement for control sample in eq 11 [−] Rf = fouling resistance [m−1] Rin = Inter layer resistance [m−1] Ro = Overall resistance [m−1] Rsu = Support layer resistance [m−1] rp = Membrane pore radius [nm] Re = Reynold number [−] Sc = Schmidt number [−] Sh = Sherwood number [−] β = Dimensionless parameter [−] Δπ = Differential osmotic pressure [Pa] ΔPeff = Effective pressure driving force [Pa] ΔP = Applied pressure [Pa] ε = Active layer porosity [−] ϵ0 = Permittivity of vacuum [8.85419 × 10−12 J−1 C2 m−1] ϵp = Pore dielectric constant [−] ζ = Surface ζ-potential [V] κ−1 = Debye length [nm] Λ = Molar conductivity [S m2 mol−1] η = Bulk viscosity [Pa s] ηapp = Apparent viscosity [Pa s] σ = Electrical conductivity of solution [S m−1] σp = Electrical conductivity of solution in the pore [S m−1] τ = Active layer tortuosity [−] REFERENCES
(1) Camargo, J. A.; Alonso, A. Ecological and toxicological effects of inorganic nitrogen pollution in aquatic ecosystems: A global assessment. Environ. Int. 2006, 32, 831. (2) Babel, S.; Kurniawan, T. A. Low-cost adsorbents for heavy metals uptake from contaminated water: A review. J. Hazard. Mater. B 2003, 97, 219. (3) Luo, Y.; Ngo, W. H. H.; Nghiem, L. D.; Hai, F. I. b.; Zhang, J.; Liang, S.; Wang, X. C. A review on the occurrence of micropollutants in the aquatic environment and their fate and removal during wastewater treatment. Sci. Tot. Environ. 2014, 473−474, 619. (4) Miralles-Cuevas, S.; Oller, I.; Sanchez Perez, J. A.; Malato, S. Removal of pharmaceuticals from MWTP effluent by nanofiltration
AUTHOR INFORMATION
Corresponding Author
*M. L. Christensen. Tel.: (+45) 9940 8464. E-mail: mlc@bio. aau.dk. Notes
The authors declare no competing financial interest. I
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Article
Industrial & Engineering Chemistry Research
(26) Madireddi, K.; Babcock, R. W.; Levine, B.; Huo, T. L.; Khan, E.; Ye, Q. F.; Neethling, J. B.; Suffet, I. H.; Stenstrom, M. K. Wastewater reclamation at Lake Arrowhead, California: An overview. Water Environ. Res. 1997, 69, 350. (27) Gherasim, C. V.; Mikulásě k, P. Influence of operating variables on the removal of heavy metal ions from aqueous solutions by nanofiltration. Desalination 2014, 343, 67. (28) Owen, G.; Bandia, M.; Howella, J. A.; Churchouse, S. J. Economic assessment of membrane processes for water and waste water treatment. J. Membr. Sci. 1995, 102, 77. (29) Pickering, K. D.; Wiesner, M. R. Cost model for low-pressure membrane filtration. J. Environ. Eng. 1993, 119, 772. (30) Xu, P.; Bellona, C.; Drewes, J. E. Fouling of nanofiltration and reverse osmosis membranes during municipal wastewater reclamation: Membrane autopsy results from pilot-scale investigations. J. Membr. Sci. 2010, 353, 111. (31) Nghiem, L.; Schafer, A.; Elimelech, M. Pharmaceutical retention mechanisms by nanofiltration membranes. Environ. Sci. Technol. 2005, 39, 7698. (32) Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.; Mariñas, B. J.; Mayes, A. M. Science and technology for water purification in the coming decades. Nature 2008, 452, 301. (33) Tsuru, T. Inorganic porous membranes for liquid phase separation. Sep. Purif. Methods 2001, 30, 191. (34) Facciotti, M.; Boffa, V.; Magnacca, G.; Jørgensen, L. B.; Kristensen, P. K.; Farsi, A.; König, K.; Christensen, M. L.; Yue, Y. Deposition of thin ultrafiltration membranes on commercial SiC microfiltration tubes. Ceram. Int. 2014, 40, 3277. (35) Farsi, A.; Boffa, V.; Qureshi, H. F.; Nijmeijer, A.; Winnubst, L.; Christensen, M. L. Modeling water flux and salt rejection of mesoporous γ-alumina and microporous organosilica membranes. J. Membr. Sci. 2014, 470, 307. (36) König, K.; Boffa, V.; Buchbjerg, B.; Farsi, A.; Christensen, M. L.; Magnacca, G.; Yue, Y. One step deposition of ultrafiltration SiC membranes on macroporous SiC supports. J. Membr. Sci. 2014, 472, 232. (37) Neamţu, M.; Grandjean, D.; Sienkiewicz, A.; Faucheur, S. L.; Slaveykova, V.; Colmenares, J. J. V.; Pulgarín, C.; de Alencastro, L. F. Degradation of eight relevant micropollutants in different water matrices by neutral photo-Fenton process under UV254 and simulated solar light irradiation A comparative study. Appl. Catal., B 2014, 158− 159, 30. (38) Pi, Y.; Schumacher, J.; Jekel, M. Decomposition of aqueous ozone in the presence of aromatic organic solutes. Water Res. 2005, 39, 83. (39) Westerhoff, P.; Aiken, G.; Amy, G.; Debroux, J. Relationships between the structure of natural organic matter and its reactivity towards molecular ozone and hydroxyl radicals. Water Res. 1999, 33, 2265. (40) Nielsen, P. H.; Raunkjær, K.; Norsker, N. H.; Jensen, N. A.; Jacobsen, T. H. Transformation of wastewater in sewer systems - A review. Water Sci. Technol. 1992, 25, 17. (41) Agirre, I.; Arias, P. L.; Castricum, H. L.; Creatore, M.; ten Elshof, J. E.; Paradis, G. G.; Ngamou, P. H.T.; van Veen, H. M.; Vente, J. F. Hybrid organosilica membranes and processes: Status and outlook. Sep. Purif. Technol. 2014, 121, 2. (42) Kanezashi, M.; Yada, K.; Yoshioka, T.; Tsuru, T. Organic− inorganic hybrid silica membranes with controlled silica network size: Preparation and gas permeation characteristics. J. Membr. Sci. 2010, 348, 310. (43) Moradi, A.; Farsi, A.; Mansouri, S. S.; Sarcheshmehpoor, M. A new approach for modeling of RO membranes using MD-SF-PF model and CFD technique. Res. Chem. Intermed. 2012, 38, 161. (44) McCleskey, R. B. Electrical conductivity of electrolytes found in natural waters from (5 to 90°C). J. Chem. Eng. Data 2011, 56, 317. (45) Rice, C. L.; Whitehead, R. Electrokinetic flow in a narrow cylindrical capillary. J. Phys. Chem. 1966, 69, 4017. (46) Van Gestel, T.; Vandecasteele, C.; Buekenhoudt, A.; Dotremont, C.; Luyten, J.; Leysen, R.; Van der Bruggen, B.; Maes,
and solar photo-Fenton using two different iron complexes at neutral pH. Water Res. 2014, 64, 23. (5) Zhou, W. H.; Cicek, N. Treatment of organic micropollutants in water and wastewater by UV-based pProcesses: A literature review. Crit. Rev. Environ. Sci. Technol. 2014, 44, 1443. (6) Wert, E. C.; Rosario-Ortiz, F. L.; Snyder, S. A. Using ultraviolet absorbance and color to assess pharmaceutical oxidation during ozonation of wastewater. Environ. Sci. Technol. 2009, 43, 4858. (7) Xu, B.; Li, D.; Li, W.; Xia, S.; Lin, Y.; Hu, C. Y.; Zhang, C.; Gao, N. Y. Measurements of dissolved organic nitrogen (DON) in water samples with nanofiltration pretreatment. Water Res. 2010, 44, 5376. (8) Lee, S.; Lueptow, R. M. Membrane rejection of nitrogen compounds. Environ. Sci. Technol. 2005, 35, 3008. (9) Juna, B.; Miyanagab, K.; Tanjib, Y.; Unno, H. Removal of nitrogenous and carbonaceous substances by a porous carrier− membrane hybrid process for wastewater treatment. Biochem. Eng. J. 2003, 14, 37. (10) Vitousek, P. M.; Aber, J. D.; Howarth, R. W.; Likens, G. E.; Matson, P. A.; Schindler, D. W. Human alteration of the global nitrogen cycle: Sources and consequences. Ecol. Appl. 1997, 7, 737. (11) National Research Council. Clean Coastal Waters: Understanding and Reducing the Effects of Nutrient Pollution; National Academic Press: Washington, DC, 2000. (12) Heidmann, I.; Calmano, W. Removal of Zn(II), Cu(II), Ni(II), Ag(I) and Cr(VI) presenting aqueous solutions by aluminum electrocoagulation. J. Hazard. Mater. 2008, 152, 934. (13) Adhoum, N.; Monser, L.; Bellakhal, N.; Belgaied, J. E. Treatment of electroplating wastewater containing Cu2+, Zn2+ and Cr(VI) by electrocoagulation. J. Hazard. Mater. 2004, 112, 207. (14) Agency for Toxic Substances and Disease Registry. Copper, September 2004. www.atsdr.cdc.gov/. (15) Marmagne, O.; Coste, C. Color removal from textile plant effluents. Am. Dyest. Rep. 1996, 15. (16) Fontanier, V.; Baig, S.; Albet, J.; Molinier, J. Comparison of conventional and catalytic ozonation for the treatment of pulp mill wastewater. Environ. Eng. Sci. 2005, 22, 127. (17) Alvares, A. B.C.; Diaper, C.; Parsons, S. A. Partial oxidation by ozone to remove recalcitrance from wastewaters − A review. Environ. Technol. 2001, 22, 409. (18) Merayo, N.; Hermosilla, D.; Blanco, L.; Cortijo, L.; Blanco, A. Assessing the application of advanced oxidation processes, and their combination with biological treatment, to effluents from pulp and paper industry. J. Hazard. Mater. 2013, 262, 420. (19) Benefield, L. D.; Morgan, J. M. Chemical precipitation. In Water Quality and Treatment; Letterman, R.D., Ed.; McGraw-Hill Inc.: New York, 1999; pp 10.1−10.57. (20) Kurniawan, T. A.; Chan, G. Y. S.; Lo, W. H.; Babel, S. Comparisons of low-cost adsorbents for treating wastewaters laden with heavy metals. Sci. Total Environ. 2006, 366, 409. (21) Kappel, C.; Kemperman, A. J. B.; Temmink, H.; Zwijnenburg, A.; Rijnaarts, H. H. M.; Nijmeijer, K. Impacts of NF concentrate recirculation on membrane performance in an integrated MBR and NF membrane process for wastewater treatment. J. Membr. Sci. 2014, 453, 359. (22) Alvarez-Ayuso, E.; Garcia-Sanchez, A.; Querol, X. Purification of metal electroplating wastewaters using zeolites. Water Res. 2003, 37, 4855. (23) Berry, L. S.; Lafayette, P. F.; Woodard, F. E. Sand filtration and activated carbon treatment of poultry process water. J.Water Pollut. Control Fed. 1976, 48, 2394. (24) Yangali-Quintanilla, V.; Sadmani, A.; McConville, M.; Kennedy, M.; Amy, G. Rejection of pharmaceutically active compounds and endocrine disrupting compounds by clean and fouled nanofiltration membranes. Water Res. 2009, 43, 2349. (25) Van der Bruggen, B.; Vandecasteele, C. Removal of pollutants from surface water and groundwater by nanofiltration: Overview of possible applications in the drinking water industry. Environ. Pollut. 2003, 122, 435. J
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Article
Industrial & Engineering Chemistry Research G. Alumina and titania multilayer membranes for nanofiltration: Preparation, characterization and chemical stability. J. Membr. Sci. 2002, 207, 73. (47) Keizer, K.; Uhlhorn, R. J. R.; Van Vuren, R. J.; Burggraaf, A. J. Gas separation mechanisms in microporous modified γ-Al2O3 membranes. J. Membr. Sci. 1988, 39, 285. (48) Kanezashi, M.; Asaeda, M. Hydrogen permeation characteristics and stability of Ni-doped silica membranes in steam at high temperature. J. Membr. Sci. 2006, 271, 86. (49) Sekulić, J.; ten Elshof, J. E.; Blank, D. H. A. A microporous titania membrane for nanofiltration and pervaporation. Adv. Mater. 2004, 16, 1546. (50) Chowdhury, S. R.; Schmuhl, R.; Keizer, K.; ten Elshof, J. E.; Blank, D. H.A. Pore size and surface chemistry effects on the transport of hydrophobic and hydrophilic solvents through mesoporous γ alumina and silica MCM-48. J. Membr. Sci. 2003, 225, 177. (51) Ju, X.; Huang, P.; Xu, N.; Shi, J. Studies on the preparation of mesoporous titania membrane by the reversed micelle method. J. Membr. Sci. 2002, 202, 63. (52) Sekulic, J.; Ten Elshof, J. E.; Blank, D. H. A. Synthesis and characterization of microporous titania membranes. J. Sol-Gel Sci. Technol. 2004, 31, 201. (53) Schattka, J. H.; Wong, E. H.M.; Antonietti, M.; Caruso, R. A. Sol−gel templating of membranes to form thick, porous titania, titania/zirconia and titania/silica films. J. Mater. Chem. 2006, 16, 1414. (54) Qureshi, H. F.; Nijmeijer, A.; Winnubst, L. Influence of sol-gel process parameters on the microstructure and performance of hybrid silica membranes. J. Membr. Sci. 2013, 446, 19. (55) ISO 6341. Water Quality - Determination of the Inhibition of the Mobility of Daphnia magna Straus (Cladocera, Crustacea)−Acute Toxicity Test; International Standards Organisation: Geneva, Switzerland, 1997. (56) ISO 10706. Water quality - Determination of Long Term Toxicity of Substances to Daphnia magna Straus (Cladocera, Crustacea); International Standards Organisation: Geneva, Switzerland, 2000. (57) ISO 11348 (part 1 and 2). Determination of the Inhibitory Effect of Water Samples on the Light Emission of Vibrio fischeri (Luminescent Bacteria Test); International Organization for Standardization: Geneva, Switzerland, 2007. (58) Roslev, P.; Bukh, A. S.; Iversen, L.; Sønderbo, H.; Iversen, N. Application of mussels as biosamplers for characterization of faecal pollution in coastal recreational waters. Water Sci. Technol. 2010, 62, 586. (59) Navalon, S.; Alvaro, M.; Garcia, H. Analysis of organic compounds in an urban wastewater treatment plant effluent. Environ. Technol. 2011, 32, 295. (60) Sekuli, J.; Magraso, A.; ten Elshof, J. E.; Blank, D. H. A. Influence of ZrO2 addition on microstructure and liquid permeability of mesoporous TiO2 membranes. Microporous Mesoporous Mater. 2004, 72, 49. (61) Bhatiaa, D.; Bourven, I.; Simon, S.; Bordas, F.; van Hullebusch, E. D.; Rossano, S.; Lens, P. N. L.; Guibaud, G. Fluorescence detection to determine proteins and humic-like substances fingerprints of exopolymeric substances (EPS) from biological sludges performed by size exclusion chromatography (SEC). Bioresour. Technol. 2013, 131, 159. (62) Ma, Q.; Liu, Y.; He, H. Synergistic effect between NO2 and SO2 in their adsorption and reaction on γ-Alumina. J. Phys. Chem. A 2008, 112, 6630. (63) Faust, S. D.; Aly, O. M. Chemistry of Water Treatment, 2nd ed.; Lewis Publishers: Boca Raton, London, New York, Washington, DC, 1999. (64) Lead, J. R.; Wilkinson, K. J.; Starchev, K.; Canonica, S.; Buffle, J. Determination of diffusion coefficients of humic substances by fluorescence correlation spectroscopy: Role of solution conditions. Environ. Sci. Technol. 2000, 34, 1365. (65) Voigt, I.; Stahn, M.; Wohner, St.; Junghans, A.; Rost, J.; Voigt, W. Integrated cleaning of coloured waste water by ceramic NF membranes. Sep. Purif. Technol. 2001, 25, 509.
(66) Marchetti, P.; Buttéa, A.; Livingston, A. G. An improved phenomenological model for prediction of solvent permeation through ceramic NF and UF membranes. J. Membr. Sci. 2012, 415, 444. (67) Lin, Y. L.; Chiou, J. H.; Lee, C. H. Effect of silica fouling on the removal of pharmaceuticals and personal care products by nanofiltration and reverse osmosis membranes. J. Hazard. Mater. 2014, 277, 102. (68) Directive 2006/7/EC of the European Parliament and of the Council of 15 February 2006 concerning the management of bathing water quality. Off. J. Eur. Union 2006, 64, 37−51.
K
DOI: 10.1021/acs.iecr.5b00064 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX