You are on page 1of 30

Subscriber access provided by Columbia Univ Libraries

Article
Highly Hydrophilic Thin-Film Composite Forward Osmosis
Membranes Functionalized with Surface-Tailored Nanoparticles
Alberto Tiraferri, Yan Kang, Emmanuel P. Giannelis, and Menachem Elimelech
ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/am301532g • Publication Date (Web): 04 Sep 2012
Downloaded from http://pubs.acs.org on September 10, 2012

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society.
1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 29 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14 Highly Hydrophilic Thin-Film Composite Forward
15
16
Osmosis Membranes Functionalized with Surface-
17 Tailored Nanoparticles
18
19
20
21
22
23
24
25 ACS Applied Materials & Interfaces
26
27 Revised: August 27, 2012
28
29
30
31
32
33
Alberto Tiraferria, Yan Kangb, Emmanuel P. Giannelisb, and Menachem
34 Elimelecha,*
35
36
37 a
38 Department of Chemical and Environmental Engineering, Yale University
39 New Haven, CT 06520-8286 (USA)
40 b
41 Department of Materials Science and Engineering, Cornell University
42 Ithaca, NY 14853 (USA)
43
44
45
46
47
48
49
50
51
52
53
54 *Tel. +1 (203) 432-2789; Fax: +1 (203) 432-4387; email: menachem.elimelech@yale.edu
55
56
57
58 1
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 2 of 29

1
2
3
4 ABSTRACT
5
6
Thin-film composite polyamide membranes are the state-of-the-art materials for membrane-
7 based water purification and desalination processes, which require both high rejection of
8
9 contaminants and high water permeabilities. However, these membranes are prone to fouling
10
11 when processing natural waters and wastewaters due to the inherent surface physicochemical
12
13 properties of polyamide. The present work demonstrates the fabrication of forward osmosis
14 polyamide membranes with optimized surface properties via facile and scalable functionalization
15
16 with fine-tuned nanoparticles. Silica nanoparticles are coated with superhydrophilic ligands
17
18 possessing functional groups that impart stability to the nanoparticles and bind irreversibly to the
19
20 native carboxyl moieties on the membrane selective layer. The tightly tethered layer of
21
22
nanoparticles tailors the surface chemistry of the novel composite membrane without altering the
23 morphology or water/solute permeabilities of the membrane selective layer. Surface
24
25 characterization and interfacial energy analysis confirm that highly hydrophilic and wettable
26
27 membrane surfaces are successfully attained. Lower intermolecular adhesion forces are
28
29 measured between the new membrane materials and model organic foulants, indicating the
30 presence of a bound hydration layer at the polyamide membrane surface that creates a barrier for
31
32 foulant adhesion.
33
34
35
36 Keywords: thin-film composite membranes, forward osmosis, superhydrophilic, membrane
37
functionalization, nanocomposite membranes, fouling, antifouling
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 2
59
60
ACS Paragon Plus Environment
Page 3 of 29 ACS Applied Materials & Interfaces

1
2
3
4 TOC Art
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 3
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 4 of 29

1
2
3
4
5 INTRODUCTION
6
7
8 Membrane-based technologies for desalination and wastewater reuse have the potential to
9 sustainably increase the supply of potable and agricultural water.1, 2
Forward osmosis, a
10
11 membrane-based separation process, is particularly attractive because it can reduce the energy of
12
13 separation.3 Polyamide thin-film composite membranes are at the core of this technology,1, 4 due
14
15 to their superior selectivity and water permeability.4, 5
However, polyamide membranes have
16
relatively high fouling propensity due to their inherent surface physicochemical properties.6, 7
17
18 Further development and implementation of processes utilizing polyamide membranes require
19
20 significant improvement in their fouling resistance. The rational optimization of polyamide
21
22 surface properties is a critical step in this endeavor.
23
24 Significant efforts have been made in membrane surface modification via post-fabrication
25
26 procedures.4, 8, 9 Recently, the use of nanomaterials has shown the potential as a novel strategy to
27
28 tailor membrane surface characteristics.10-12 Nanoparticles can be fine-tuned in terms of size,
29 shape, and surface chemistry to achieve desired characteristics. Well-designed nanoparticles can
30
31 have strong interactions with the membrane surface to ensure irreversible functionalization and
32
33 retain the targeted property during operation. The effective combination of such materials with
34
35 polymeric membranes would consist of non-depleting and scalable surface functionalizations
36
37
that do not affect other crucial transport parameters of the membrane.
38
39 Recent studies have made use of antimicrobial nanoparticles with the goal to impart biocidal
40
41
properties to polyamide membrane and control their biofouling.11, 12
These studies have
42 suggested ways to permanently tether nanoparticles by exploiting the native functional groups of
43
44 polyamide. In addition, it was shown that controlling the surface density and uniform
45
46 distribution of the nanoparticle coating is especially important to concentrate the nanoparticle
47
48 activity at the membrane surface.12
49
50 Inactivation of microorganisms that attach to the membrane would delay the onset of biofilm
51
52 formation.13-15 However, the primary attachment mechanism of microorganisms involves the
53
secretion of protein-based adhesives.16-18 Additionally, many other organic molecules are
54
55 present in feedstreams and contribute significantly to the decrease in process performance due to
56
57
58 4
59
60
ACS Paragon Plus Environment
Page 5 of 29 ACS Applied Materials & Interfaces

1
2
3
fouling. Therefore, development of surface functionalizations that decrease the adhesion or
4
5 increase the release of adsorbed organic molecules is of paramount importance.1, 17
6
7
8
Wettability and hydrophilicity of the membrane material play a crucial role in controlling
9 fouling resistance and release of adsorbed foulants, as these parameters are directly related to the
10
11 material surface tension.1, 17, 19 Specifically, by increasing the number of hydrogen bonding sites
12
13 at the surface, the interfacial acid-base forces are maximized, thereby allowing the formation of
14
15 an interfacial layer of tightly bonded water molecules that are highly oriented and have slow
16 dynamics. Because the displacement of water molecules involves work that increases the free
17
18 energy of the system, this hydration layer provides a repulsive barrier against the adsorption of
19
20 foulants.20 Studies have shown that increasing the membrane hydrophilicity can hinder fouling
21
22 by microorganisms,21 proteins,22-24 lipopolysaccharides,23 inorganic colloids,25 alginate,26 and
23
24
natural organic matter.26
25
26 The present study demonstrates the fabrication of highly hydrophilic thin-film composite
27
28
polyamide forward osmosis membranes by surface functionalization with tailored nanoparticles.
29 The proposed surface functionalization procedure is remarkably simple and effective, and
30
31 follows the steps illustrated in Scheme 1. Silica nanoparticles (Step A) are surface-coated with
32
33 superhydrophilic cationic ligands (Step B) to create a stable nanoparticle suspension. The
34
35 ligands are terminated with either quaternary ammonium or amine functional groups (Step C), to
36 stabilize the nanoparticles and to provide anchor sites for tethering the nanoparticles to the
37
38 membranes. A dip-coating protocol is performed, during which the nanoparticles strongly bind
39
40 to the native carboxyls of hand-cast polyamide forward osmosis membranes (Step D). The
41
42 newly fabricated surfaces (Step E) are extensively characterized and their physicochemical
43
44
properties as well as their interfacial energies are investigated. The new hydrophilic membranes
45 have the potential to significantly improve membrane performance by reducing and delaying
46
47 fouling.
48
49
50
51
MATERIALS AND METHODS
52
53 Fabrication of the Membranes and Characterization of their Transport
54
55 Properties. TFC forward osmosis membranes were prepared by interfacial polymerization of
56
57
58 5
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 6 of 29

1
2
3
polyamide onto hand-cast support membranes. The support membranes were fabricated by
4
5 nonsolvent (water) induced phase separation of a solution of 9 wt% polysulfone (PSf, Mn: 22,000
6
7 Da) dissolved in N-N-dimethylformamide (DMF, anhydrous, 99.8%), following the procedure
8
9 outlined in our previous publication.27 The polyamide active layer was then formed on top of the
10
11
PSf support membranes via reaction between 1,3-phenylenediamine (MPD, >99%) and 1,3,5-
12 benzenetricarbonyl trichloride (TMC, 98%) dissolved in Isopar-G (Univar, Redmond, WA), as
13
14 described in our previous publication.12
15
16
17 The transport properties of the control and functionalized forward osmosis membranes were
18
19
tested using a cross-flow membrane filtration system in forward osmosis and pressure-retarded
20 osmosis configurations.28, 29
These measurements allowed for the determination of the pure
21
22 water permeability of the membrane active layer, A, the salt (NaCl) permeability of the
23
24 membrane active layer, B, and the structural parameter of the membrane support layer, S, as 2.46
25
26 ± 1.19 L m-2h-1bar-1, 1.70 ± 1.18 L m-2h-1, and 537 ± 182 µm, respectively (Table S4 of the
27 Supporting Information). Details of the characterization protocol can be found elsewhere.30
28
29
30 Fabrication and Characterization of the Superhydrophilic Nanoparticles.
31
32 Superhydrophilic nanoparticles were fabricated by surface functionalization of silica
33
34
nanoparticles (Ludox HS-30, 30%, Sigma Aldrich) with two different ligands (Scheme 1, steps
35 A-B-C). In the first instance, 6 g of silica nanoparticles were dispersed in 30 mL of deionized
36
37 water and the suspension was sonicated for 30 min. The obtained dispersion was vigorously
38
39 stirred with freshly prepared silane solution containing 2.1 g of (3-aminopropyl)trimethoxysilane
40
41 (—NH3+/NH2, 97%, Sigma-Aldrich 281778) dissolved in 24 mL of water. For the second
42
43
functionalization, 6 g of silica nanoparticles were suspended in 54 mL of deionized water and
44 sonicated for 30 min. Then, 6.4 g of N-trimethoxysilylpropyl-N,N,N-trimethylammonium
45
46 chloride (—N(CH3)3+, 50 wt%, Gelest SIT8415.0) were added to the dispersion under vigorous
47
48 stirring. Both procedures were followed by pH adjustment to ~5 and a heating step at 60 °C for
49
50 18 hr. Finally, the suspensions were dialyzed in deionized water using SnakeSkin tubing (7k
51
52
MWCO, Pierce) for 48 hours.
53
54 Dynamic light scattering (DLS) experiments were performed to determine the effective
55
56 hydrodynamic diameters of the functionalized nanoparticles using a multi-detector light
57
58 6
59
60
ACS Paragon Plus Environment
Page 7 of 29 ACS Applied Materials & Interfaces

1
2
3
scattering unit (ALV-5000, Langen, Germany) following the procedure outlined in our previous
4
5 publication.31 The electrophoretic mobility of the particles was determined by a Zetasizer Nano-
6
7 Z (Malvern Instruments, Worcestershire, U.K.) in deionized water at three different pH values of
8
9 5, 6, and 7. For thermogravimetric analysis (TGA) (Exstar TG/DTA 6200, Seiko Instruments
10
11
Inc., Torrance, CA), the nanoparticle solution was freeze-dried and TGA was performed from 40
12 to 600 °C at a heating rate of 20 °C/min. Transmission Electron Microscopy (TEM)
13
14 micrographs of the nanoparticles were acquired using a Tecnai T12 apparatus operating at 120
15
16 keV (FEI, Eindhoven, The Netherlands).
17
18
19
Membrane Functionalization and Characterization. The density of carboxyl
20 functional groups at the surface of polyamide membranes was evaluated by binding and elution
21
22 of toluidine blue O dye (TBO), as described in our recent publication.32 Carboxyl moieties were
23
24 exploited to irreversibly bind the functionalized silica nanoparticles to the membranes, following
25
26 a simple dip coating protocol (Scheme 1, steps D-E). Briefly, the polyamide membranes were
27 immersed into the nanoparticle suspension for 16 hr at room temperature (23 °C), with only the
28
29 membrane active layer side accessible for contact with the suspension. The pH of the
30
31 suspensions was adjusted to between 6.4 and 7.4 before the dip-coating step. In the case of
32
33 membrane functionalization with nanoparticles coated with amine-terminated ligands, the
34
tethering procedure was preceded by contact of the polyamide layer with a solution of ~2 mM N-
35
36 (3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC, 98%) and ~5 mM N-
37
38 hydroxysuccinimide (NHS, 98%) for 15 min. The polyamide surface treatment with EDC and
39
40 NHS converts the native carboxylate groups of the polyamide surface into intermediate amine-
41
42 reactive esters33 for crosslinking with the amine functional groups at the nanoparticle surface.
43
44 The elemental composition of the membrane surface was analyzed by x-ray photoelectron
45
46 spectroscopy (XPS, SSX-100 UHV, Surface Science Instruments). The sample was irradiated
47 with a beam of monochromatic Al K-alpha X-rays with energy of 1.486 keV. Attenuated Total
48
49 Reflectance (ATR-IR, ThermoScientific Nicolet 6700) was performed using a germanium crystal
50
51 on desiccator-dried samples. Membrane surface morphology was investigated by scanning
52
53 electron microscopy (SEM, LEO 1550 FESEM). Before imaging, membranes were sputter
54
55
coated with a layer of carbon (BTT-IV, Denton Vacuum, LLC, Moorestown, NJ). Membrane
56 surface roughness was analyzed using a Multimode AFM (Veeco Metrology Group, Santa
57
58 7
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 8 of 29

1
2
3
Barbara, CA) in tapping mode. Symmetric silicon probes with 30-nm-thick back side aluminum
4
5 coating were employed (Tap300A, Bruker Nano, Inc., Camarillo, CA). The probe had a spring
6
7 constant of 40 N/m, resonance frequency of 300 kHz, tip radius of 8 ± 4 nm, and cantilever
8
9 length of 125 ± 10 µm. Air-dried membranes were scanned in air at 12 randomly selected scan
10
11 positions.
12
13 Surface wettability was evaluated from contact angle measurements of deionized water using
14
15 the sessile drop method (VCA Video Contact Angle System, AST Products, Billerica, MA). The
16
17
system is equipped with software to determine the left and right contact angles (VCA Optima
18 XE). To account for variations between different measurements on the same surface, at least 4
19
20 desiccator-dried samples from separately cast and functionalized membranes were tested on a
21
22 minimum of six random locations, and the data were averaged. The relative wettability of the
23
24 membranes was evaluated by calculating the membrane-liquid interfacial free energy
25
 cos θ  , where θ is the average contact angle and γ is the pure water surface
26 as −∆G
=ML γ L 1 +  L
27  r 
28
29 tension (72.8 mJ/m2 at 25 °C), and r is the roughness area ratio (i.e., the ratio of actual surface
30
31
area for a rough surface to the planar area, r = 1 + SAD , with SAD being the surface area
32 difference parameter).34, 35
Contact angles of deionized water were also used as a proxy to
33
34 confirm the irreversibility of the nanoparticle-membrane bonds with functionalized membrane
35
36 surfaces, after these were subjected to chemical or physical stress. Chemical stress was applied
37
38 by contacting the functionalized surfaces for 15 min with a pH 2 solution (HCl), a pH 12 solution
39
(NaOH), or a 0.6 M NaCl solution approximating the ionic strength of typical seawater, followed
40
41 by a thorough rinse with deionized water. Physical stress was exerted by immersing the
42
43 membranes in a sonicating water bath (Fisher Scientific F60) for 7 min. XPS spectra and SEM
44
45 images were also re-evaluated after each of these steps to confirm the presence and extent of
46
47
particle functionalization and assess the irreversibility of the functionalization.
48
49 Additional measurements of contact angles of glycerol (≥99%) and diiodomethane (≥99%)
50
51 were used to calculate the Lifshitz-van der Waals (γLW), electron donor (γ-), and electron acceptor
52 (γ+) components of the membrane surface tension before and after functionalization.36-38 In these
53
54 calculations, the excess surface area due to roughness is accounted for by incorporation of the
55
56 roughness area ratio, r, which was defined earlier. The total surface energy of the membrane
57
58 8
59
60
ACS Paragon Plus Environment
Page 9 of 29 ACS Applied Materials & Interfaces

1
2
3
surfaces is defined as the sum of the surface tension due to Lifshitz-van der Waals and Lewis
4
5
6 acid-base components, γ TOT
= γ LW + γ AB , where γ AB = 2 γ +γ − .36, 37 From the membrane and the
7
8 water components of the surface tension, it is possible to calculate the total interfacial free
9
10 energy of cohesion of membrane interfaces immersed in water, ΔGMLM(TOT), which is often
11
termed “hydrophilicity”.36-38 A higher value of the free energy is obtained if the membrane is
12
13 non-cohesive, or more hydrophilic, when immersed in water.
14
15
16
The zeta potential of the membrane surface before and after functionalization was measured
17 in an asymmetric clamping cell using a streaming potential analyzer (EKA, Brookhaven
18
19 Instruments, Holtsville, NY). Measurements were performed while alternating flow direction of
20
21 a 1 mM KCl solution, and by varying the pH of the solution by adding appropriate amounts of
22
23 HCl or KOH. Four separately cast and functionalized membranes were evaluated. A detailed
24 experimental procedure and the method used to calculate the zeta potential from the measured
25
26 streaming potential are given elsewhere.39
27
28
29 AFM Interaction Forces. Atomic force microscopy (AFM) was used to measure the
30
31 adhesive force between representative foulants in the bulk solution and the membrane by
32 adapting the procedures described by Li and Elimelech.40 The force measurements were
33
34 performed in a fluid cell utilizing a particle probe, modified from a commercialized SiN AFM
35
36 probe (Veeco Metrology Group, Santa Barbara, CA). A carboxylate modified latex (CML)
37
38 particle (Interfacial Dynamics Corp., Portland, OR) with a diameter of 4.0 µm was attached to
39
40 the tipless SiN cantilever using Norland Optical adhesive (Norland Products, Inc., Cranbury,
41 NJ). The particle probe was cured under UV light for 30 min. The CML-modified probe was
42
43 immersed in 2000 mg/L model organic foulant solution, namely alginate or bovine serum
44
45 albumin (BSA), for at least 16 hr at 4 °C to prevent organic degradation. The AFM adhesion
46
47 force measurements were performed in a fluid cell. The ionic composition of the test solutions
48
injected into the fluid cell was representative of a typical wastewater effluent (0.45 mM KH2PO4,
49
50 9.20 mM NaCl, 0.61 mM MgSO4, 0.5 NaHCO3, 0.5 mM CaCl2, and 0.935 mM NH4Cl).13 The
51
52 pH of the test solution was adjusted to 7.4 prior to injection. The membrane was equilibrated
53
54 with the test solution for 30 to 45 min before force measurements were performed. The force
55
56
measurements were conducted at five different locations, and at least 25 measurements were
57
58 9
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 10 of 29

1
2
3
taken at each location. Data obtained from the retracting force curves were processed and
4
5 converted to obtain the force versus surface-to-surface separation curves.
6
7
8 RESULTS AND DISCUSSION
9
10 Properties of the Nanoparticles are Fine-Tuned for Membrane Functionalization.
11
12 Silica nanoparticles were used because their surface chemistry can be readily fine-tuned, thereby
13
14 facilitating the attainment of target hydrophilic properties and enabling control of the interaction
15
with the membrane surface. Two different ligands were employed to functionalize the
16
17 nanoparticle surface. Nanoparticles treated with N-trimethoxysilylpropyl-N,N,N-
18
19 trimethylammonium chloride carry quaternary ammonium groups and are hereafter designated as
20
21 —N(CH3)3+ nanoparticles. The second treatment using (3-aminopropyl)trimethoxysilane
22
23 produced nanoparticles with amine surface functionalities that are henceforth referred to as —
24
25
NH2/NH3+ nanoparticles.
26
27 Upon surface functionalization, the presence of ammonium or amine groups rendered the
28
29 functionalized nanoparticles positively charged, as confirmed by measurements of their
30 electrophoretic mobility (Figure 1). The charge of—N(CH3)3+ nanoparticles is not significantly
31
32 affected by solution pH, while the charging behavior of the—NH2/NH3+ nanoparticles is
33
34 dependent on solution pH through protonation/deprotonation.
35
36 The starting bare silica nanoparticles had a hydrodynamic radius of approximately 7 nm as
37
38 observed by DLS measurements. The measured radius in deionized (DI) water increased to ~8
39
40 and ~19 nm for the —N(CH3)3+ and —NH2/NH3+ functionalizations, respectively (Figure 1,
41
42 table). While the small increase in diameter for the quarternary ammonium-functionalized
43
44 nanoparticles is attributed to the presence of a hydration layer bound to the hydrophilic surface
45
46
ligands, the increase in size of the amine nanoparticles was likely due to slight aggregation.
47 TEM imaging confirmed that the size of both types of functionalized nanoparticles was
48
49 comparable to that of the bare silica nanoparticles (not shown). This observation substantiates
50
51 our hypothesis that the —NH2/NH3+ nanoparticles undergo aggregation in aqueous solution. No
52
53 change in diameter was observed by DLS within 45 minutes of measurement for both
54
functionalized nanoparticle types, suggesting that aggregation occurred immediately upon
55
56 dispersion of the particle in solution. Overall, the positively-charged surface groups increased
57
58 10
59
60
ACS Paragon Plus Environment
Page 11 of 29 ACS Applied Materials & Interfaces

1
2
3
the electrostatic repulsion between functionalized nanoparticles, thwarting their aggregation in
4
5 aqueous solution.
6
7
8
In the presence of electrolytes in solution, DLS data demonstrated an increase in
9 hydrodynamic size for all nanoparticles (Figure 1, table). This phenomenon can be due to slight
10
11 aggregation and/or to the adsorption of highly hydrated multivalent counterions onto the charged
12
13 and hydrophilic particle surface. This mechanism could further enhance the structuring of the
14
15 water molecules at the solid-liquid interface, resulting in a larger measured hydrodynamic
16 diameter by DLS.41-43
17
18
19 The presence of organic ligands on the surface of the functionalized nanoparticles was
20
confirmed by TGA measurement (Figure 1C-D-E). TGA data showed the appearance and
21
22 amplification of two thermal degradation peaks (~250 and ~400 °C) for the functionalized
23
24 nanoparticles. These peaks may be associated with thermo-oxidation of the alkyl chains of the
25
26 surface ligands and possibly to the volatilization of some excess coupling agents used during
27
28
particle functionalization. The production of a larger amount of volatile degradation products
29 translated into a smaller percentage of sample recovery at the end of the heating cycle compared
30
31 to the bare silica nanoparticles.
32
33
34 Nanoparticles are Irreversibly Bound to the Membrane Surface after
35
36
Functionalization. All polyamide membranes fabricated via interfacial polymerization of
37 TMC and MPD inherently possess an outer layer of relatively high, negative fixed charges
38
39 resulting from incomplete reaction and hydrolysis of the TMC acyl chlorides into carboxyls.32, 44
40
41 The surface density of carboxylic groups of the membranes used in this study was measured by
42
43 TBO 19 ± 4 charges/nm2 of planar area. The positively charged groups at the nanoparticle
44
45
surface ensure durable adhesion to the membrane surface via strong interaction with the native
46 polyamide moieties, thus securing the nanoparticles at this interface. Specifically, the
47
48 membrane−particle tethering occurred here primarily via electrostatic attraction.45 In addition,
49
50 the functionalization with —NH2/NH3+ nanoparticles was performed in the presence of
51
52 crosslinking agents EDC and NHS to facilitate the formation of covalent amide bonds between
53
54 the nanoparticle amine groups and the membrane carboxyls.12 The functionalized membranes
55 are hereafter designated as —N(CH3)3+ or —NH2/NH3+ membranes.
56
57
58 11
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 12 of 29

1
2
3
XPS data of the membrane surfaces evaluated before and after functionalization are
4
5 presented in Figure 2. The energy peaks observed for the polyamide surface (black) are
6
7 attributed to carbon, oxygen, and nitrogen (Figure 2A) among which carbon was the most
8
9 abundant element (Figure 2B), consistent with the chemistry of the membrane active layer. The
10
11
spectra related to the functionalized surfaces (blue and red) showed the appearance of energy
12 peaks associated with silicon (Figure 2A), which confirm the presence of the silica-based
13
14 nanoparticles at these surfaces. Because XPS analyzes only the superficial portion of the
15
16 membrane, oxygen was observed to be the predominant element, followed by carbon and silicon
17
18 (Figure 2C-D), according to the composition of the functionalized silica nanomaterial. ATR-IR
19 spectra showed the emergence of a shoulder and an increase in absorbance around 1060 to 1100
20
21 cm-1 (Supporting Information Figure S1), which is attributed to the stretching mode of Si-O-Si
22
23 bonds.46 This observation further confirms the presence of silanized SiO2 nanoparticles at the
24
25 membrane surface.
26
27 Physicochemical Properties of the Membrane Surface. Figure 3 presents the
28
29 pH−dependent zeta potential of the membrane surfaces before and after functionalization. The
30
31 zeta potential was measured over the pH range of 4 to 9 for at least 4 separately cast and
32
functionalized membrane samples. Knowledge of the membrane surface zeta potential and of
33
34 the type and density of exposed charges is crucial because these parameters greatly influence the
35
36 membrane fouling behavior.40, 47
37
38 The results with the control membranes were in accordance with the protonation behavior of
39
40 polyamide functional groups. At very low pH, the unreacted amine groups of MPD are
41
42 protonated while carboxylic groups are uncharged, resulting in an overall positive potential
43
44 (Figure 3A). As the pH increased above the pKa of the polyamide carboxyl groups,48 these
45 predominant acidic groups deprotonated, thus imparting a negative and largely constant zeta
46
47 potential to the surface.
48
49
The zeta potential behavior of the functionalized membranes was consistent with the
50
51 functionalities present at both the nanoparticle and the membrane surface. The —N(CH3)3+
52
53 nanoparticles are positively charged at all pH values and interact with the membrane carboxylic
54
55 moieties via electrostatic attraction. Therefore, the zeta potential of the membranes was highly
56
57
58 12
59
60
ACS Paragon Plus Environment
Page 13 of 29 ACS Applied Materials & Interfaces

1
2
3
positive at low pH, where carboxyl groups are uncharged, and became progressively more
4
5 negative as the carboxylic groups deprotonated (Figure 3B). The overall zeta potential was close
6
7 to zero around the pH range of 7 to 8, which is the characteristic pH of natural waters and
8
9 wastewater effluents in membrane separation processes.
10
11 Nanoparticles functionalized with —NH2/NH3+ ligands are assumed to preferentially form
12
13 amide bonds with the membrane carboxylic groups, thus effectively neutralizing many of the
14
15 charges present on both reacting surfaces. As a result, the measured zeta potential values of the
16
17
—NH2/NH3+ membranes were of lower magnitude compared to those of the —N(CH3)3+
18 membranes and exhibited a wider near-zero potential region, between approximately pH 6 and 8
19
20 (Figure 3C). The zeta potential results provide indirect evidence for the presence of nanoparticles
21
22 at the surface of the functionalized membranes and for the type of particle−membrane
23
24 interaction.
25
26 The membrane surface morphology before and after functionalization was analyzed by SEM
27
28 and AFM (Figure 4). The representative topographic image (Figure 4G) and SEM surface
29
30 micrographs (Figure 4A-B) of a control polyamide membrane showed a uniform ridge-and-
31
valley morphology, which is typical of polyamide thin films formed by interfacial
32
33 condensation.1, 4
The characteristic surface roughness parameters of the membranes were
34
35 measured by tapping mode AFM. The untreated polyamide surfaces had a RMS of 129 ± 40 nm,
36
37 an average roughness, Ra, of 102 ± 39 nm, a maximum roughness, Rmax, of 850 ± 30 nm, and a
38
39 surface area difference, SAD, of 23 ± 10% (Figure 4H). These values are comparable to those
40
41
reported for similar materials.49
42
43 The high magnification SEM micrographs in Figures 4B-D-F, imaged at the surface of the
44
45 membranes after functionalization, showed that the ridge-and-valley features of the
46 functionalized surfaces were overlaid by a layer of nanoparticles. The nanoparticle size
47
48 correlates well with the radius measured by DLS experiments for each respective type of surface
49
50 functionality. The low magnification SEM micrographs presented in Figures 4A-C-E suggest
51
52 that the overall morphology of the membrane surface was not significantly affected after
53
functionalization, as the ridge-and-valley features were visible and comparable to those observed
54
55 for the control polyamide surface. This observation suggests that the nanoparticle coating
56
57
58 13
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 14 of 29

1
2
3
consisted of a mono- or multi-layer, with a total thickness small relative to the roughness of the
4
5 membrane active layer.
6
7
8
The surface roughness measurements of functionalized membranes (Figure 4H) indicated a
9 reduction in surface roughness due to the presence of nanoparticles, although it was not
10
11 sufficient to alter the overall surface morphology, consistent with SEM analysis. The
12
13 nanoparticles are likely to deposit preferentially within the valley-like regions of the polyamide
14
15 surface, thus flattening the overall surface. This flattening was more pronounced for the
16
relatively larger —NH2/NH3+ nanoparticles, which produced a more significant effect in
17
18 decreasing the membrane SAD (Figure 4H and Table 1).
19
20
21 Nanoparticles Render the Membrane Highly Hydrophilic. Figure 5 presents the
22
23 average contact angles of DI water at the surface of control (polyamide) and functionalized
24
membranes before (solid bar) and after (hollow bars) they were subjected to chemical and
25
26 physical stresses. The untreated polyamide membranes had a contact angle of 104 ± 16°, larger
27
28 than what typically observed for commercial polyamide membranes (Figure 5, Table 1). For
29
30 rough surfaces, such as those of our hand-cast samples, larger contact angles than expected from
31
32
the chemistry of the material might result from air trapped between the solid surface rugosities
33 and the liquid droplet.50 The digital picture (Figure 5A) shows a representative profile of a water
34
35 droplet on the hydrophobic polyamide surface. The presence of nanoparticles on the surface
36
37 functionalized membranes had a dramatic effect on the conformation of water droplets at the
38
39 solid-liquid interface, yielding contact angles of ~10° for the —N(CH3)3+ membranes and ~20°
40
for the —NH2/NH3+ membranes (Table 1), only in part attributable to the decrease in roughness
41
42 measured on the functionalized surfaces (Figure 4). Representative pictures of water droplet
43
44 profiles for the two functionalized surfaces are presented on the right of Figure 5B.
45
46 Contact angle measurements were also used as a proxy to appraise the reversibility of the
47
48 interaction between nanoparticles and membrane surfaces. Chemical or physical stresses
49
50 considerably harsher than typical operational conditions were applied to the functionalized
51
52 membranes and the conformation of water droplets was then re-evaluated. The contact angles
53 did not significantly change compared to membranes analyzed immediately after modification
54
55 (Figure 5A and B), suggesting that the nanoparticle-membrane bonds were sufficiently strong to
56
57
58 14
59
60
ACS Paragon Plus Environment
Page 15 of 29 ACS Applied Materials & Interfaces

1
2
3
render the surface functionalization irreversible. XPS and SEM analyses were also performed
4
5 subsequent to the stress protocol and showed no significant difference compared to the results
6
7 obtained on the functionalized membranes not subjected to stresses (Supporting Information
8
9 Figure S2). The confirmed strength of the nanoparticle-membrane interaction suggests minimal
10
11
detachment of the nanoparticles from the surface during typical membrane operational
12 conditions, thus ensuring long-term functionalization.
13
14
15 The surface tensions and interfacial free energies of the membranes were calculated from
16 contact angle measurements with two polar liquids, water and glycerol, and an apolar liquid,
17
18 diiodomethane (Table 1). The polyamide control membrane had a relatively low surface energy
19
20 ( γ TOT = 37.8 mJ/m2), almost exclusively resulting from van der Waals forces. As a result, the
21
22 polyamide surface was found to be relatively wetting ( −∆GML = 57.9 mJ/m2), but hydrophobic
23
24 ( ∆GMLM =
−97.9 mJ/m2) when immersed in deionized water (Figure 6). These results are
25
26 consistent with data obtained from other polyamide membranes.38, 51
27
28
29 The surface properties of the membranes changed dramatically after functionalization with
30 superhydrophilic nanoparticles. Both the Lifshitz-van der Waals and the acid-base components
31
32 of surface tensions increased. In particular, the electron donor parameter was responsible for the
33
34 nearly monopolar functionality of the surface (Table 1), consistent with the properties of the
35
36 ligands coating the nanoparticle surface.52, 53
The high density of electron donor sites at the
37
38
surface of the functionalized membranes promotes hydrogen bonding interactions with water
39 molecules.54 This, in turn, resulted in a significant increase in calculated membrane wettability
40
41 and in a conversion of the surface interfacial free energy of cohesion to positive values, i.e.
42
43 hydrophilic properties (Figure 6A). The values of hydrophilicity estimated for our
44
45 functionalized membranes are the highest reported so far in the literature for similar materials.
46 The high interfacial free energy was accompanied by a relatively large value of surface energy
47
48 (Figure 6B). The strong hydration layer of the near superhydrophilic surface resists the
49
50 adsorption of molecules and particles to the membrane surface, thus increasing its anti-fouling
51
52 resistance.17
53
54 Hydrophilic Membranes Have Lower Interaction Forces with Organic Foulants.
55
56 The rationale for creating highly hydrophilic membranes for water separation technologies is to
57
58 15
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 16 of 29

1
2
3
impart fouling resistance. By maximizing the interfacial acid-base forces between surfaces and
4
5 the adherent water, we form a layer of tightly bonded water molecules that act as a short-range
6
7 barrier against the adhesion of foulants.20 Atomic force microscopy (AFM) has been applied in
8
9 membrane fouling/cleaning research to quantify intermolecular forces when foulants approach
10
11
the investigated surface within the contact limit.6, 20, 40, 55, 56 In this study, we investigated the
12 interaction forces between model foulants adsorbed on a colloidal probe, namely alginate and
13
14 BSA, and our membranes (Figure 7). Representative adhesion (pull-off) curves obtained during
15
16 the retraction of the fouled tip from the membrane surface are presented. We also report the
17
18 average, minimum, and maximum values of adhesion forces calculated from a statistically
19 significant number of retracting force-distance curves analyzed in 5 randomly selected spots on
20
21 each membrane sample.
22
23
24
AFM results showed that the attractive energy well between model foulants and control
25 polyamide membranes was deeper than that observed using functionalized, hydrophilic
26
27 membranes (Figure 7A and B). The resulting distribution of foulant-membrane intermolecular
28
29 forces was also statistically more negative (i.e. more attractive) for the control polyamide
30
31 membranes (Supporting Information Figure S3). Several force-distance curves measured on —
32
N(CH3)3+ membranes did not show an attractive energy well but only repulsive forces, indicating
33
34 no foulant adhesion to the membrane due to a barrier to adhesion.57 This behavior was not
35
36 observed for control polyamide membranes on which all AFM foulant probe engagements
37
38 resulted in an attractive force, often exceeding -3 mN/m for both foulant molecules. These
39
40
results are consistent with previous observations showing lower attractive forces on hydrophilic
41 surfaces,20 and indicate the attainment of highly hydrophilic surfaces with potentially lower
42
43 fouling propensity. The antifouling behavior and antifouling mechanisms of our highly
44
45 hydrophilic membranes were further confirmed in cross-flow experiments in forward osmosis
46
47 and reverse osmosis operation modes, using feed solutions of BSA or alginate as described in our
48
related publication.30
49
50
51
52 CONCLUSION
53
54 We fabricated forward osmosis membranes with near superhydrophilic surface properties that
55
56 could significantly reduce fouling. The surface of silica nanoparticles was functionalized with
57
58 16
59
60
ACS Paragon Plus Environment
Page 17 of 29 ACS Applied Materials & Interfaces

1
2
3
superhydrophilic ligands possessing quaternary ammonium or amine moieties. A simple dip-
4
5 coating technique was utilized to irreversibly bind the nanoparticles to the native carboxylic
6
7 groups of polyamide forward osmosis membranes. The functionalization produced a uniform
8
9 layer of nanoparticles on the polyamide film rendering the membrane surface highly wettable
10
11
and hydrophilic. The post-fabrication route to functionalization ensures that the productivity and
12 the rejection performance of the membranes are maintained. Using atomic force microscopy, we
13
14 measured significantly lower adhesion forces between model organic foulants and the
15
16 hydrophilic surfaces compared to unmodified polyamide membranes. These observations are
17
18 significant because lower foulant-membrane adhesion has been shown to correlate well with
19 increased membrane fouling resistance.
20
21
22
23
24 ASSOCIATED CONTENT
25
26
27 Supporting Information. ATR-IR analysis of the membrane surfaces confirming presence
28
29 of silica nanoparticles after membrane functionalization (Figure S1); XPS and SEM analyses of
30
the membrane surface performed after subjecting the membrane surface to stress, confirming the
31
32 irreversibility of surface functionalization (Figure S2); statistics of foulant-membrane interaction
33
34 forces measured by AFM (Figure S3); summary of forward osmosis performance testing
35
36 conditions and results (Table S4).
37
38
39
40
41 AUTHOR INFORMATION
42
43
44 Corresponding Author. *Email: menachem.elimelech@yale.edu; phone: (203) 432-2789
45
46
47
48
49
50
ACKNOWLEDGMENTS
51
52 This publication is based on work supported by Award No. KUS-C1-018-02, made by King
53
54 Abdullah University of Science and Technology (KAUST). We also acknowledge the NWRI-
55 AMTA Fellowship for Membrane Technology, awarded to Alberto Tiraferri.
56
57
58 17
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 18 of 29

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 18
59
60
ACS Paragon Plus Environment
Page 19 of 29 ACS Applied Materials & Interfaces

1
2
3
4
5 REFERENCES
6
7
8 (1) Elimelech, M.; Phillip, W. A., Science 2011, 333 (6043), 712.
9 (2) McGinnis, R. L.; Elimelech, M., Environ. Sci. Technol. 2008, 42 (23), 8625.
10 (3) McGinnis, R. L.; Elimelech, M., Desalination 2007, 207 (1-3), 370.
11
(4) Petersen, R. J., J. Membr. Sci. 1993, 83 (1), 81.
12
13 (5) Geise, G. M.; Park, H. B.; Sagle, A. C.; Freeman, B. D.; McGrath, J. E., J. Membr. Sci.
14 2011, 369 (1-2), 130.
15 (6) Ang, W. S.; Tiraferri, A.; Chen, K. L.; Elimelech, M., J. Membr. Sci. 2011, 376 (1-2),
16 196.
17 (7) Mi, B. X.; Elimelech, M., J. Membr. Sci. 2010, 348 (1-2), 337.
18
19
(8) Rana, D.; Matsuura, T., Chem. Rev. 2010, 110 (4), 2448.
20 (9) Kang, G. D.; Cao, Y. M., Water Res. 2012, 46 (3), 584.
21 (10) Kim, S. H.; Kwak, S. Y.; Sohn, B. H.; Park, T. H., J. Membr. Sci. 2003, 211 (1), 157.
22 (11) Mo, J.; Son, S. H.; Jegal, J.; Kim, J.; Lee, Y. H., J. Appl. Polym. Sci. 2007, 105 (3),
23 1267.
24
(12) Tiraferri, A.; Vecitis, C. D.; Elimelech, M., Acs Applied Materials & Interfaces 2011, 3
25
26 (8), 2869.
27 (13) Herzberg, M.; Elimelech, M., J. Membr. Sci. 2007, 295 (1-2), 11.
28 (14) De Kwaadsteniet, M.; Botes, M.; Cloete, T. E., Nano 2011, 6 (5), 395.
29 (15) Upadhyayula, V. K. K.; Gadhamshetty, V., Biotechnol. Adv. 2010, 28 (6), 802.
30 (16) Stanley, M. S.; Callow, M. E.; Callow, J. A., Planta 1999, 210 (1), 61.
31
32
(17) Callow, J. A.; Callow, M. E., Nature Communications 2011, 2.
33 (18) Herzberg, M.; Kang, S.; Elimelech, M., Environ. Sci. Technol. 2009, 43 (12), 4393.
34 (19) Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.; Marinas, B. J.; Mayes,
35 A. M., Nature 2008, 452 (7185), 301.
36 (20) Morra, M., Journal of Biomaterials Science-Polymer Edition 2000, 11 (6), 547.
37
(21) Finlay, J. A.; Callow, M. E.; Ista, L. K.; Lopez, G. P.; Callow, J. A., Integr. Comp. Biol.
38
39 2002, 42 (6), 1116.
40 (22) Wavhal, D. S.; Fisher, E. R., Langmuir 2003, 19 (1), 79.
41 (23) Gudipati, C. S.; Finlay, J. A.; Callow, J. A.; Callow, M. E.; Wooley, K. L., Langmuir
42 2005, 21 (7), 3044.
43 (24) Ostuni, E.; Chapman, R. G.; Holmlin, R. E.; Takayama, S.; Whitesides, G. M.,
44
45
Langmuir 2001, 17 (18), 5605.
46 (25) Brant, J. A.; Childress, A. E., J. Membr. Sci. 2004, 241 (2), 235.
47 (26) Ang, W. S.; Lee, S. Y.; Elimelech, M., J. Membr. Sci. 2006, 272 (1-2), 198.
48 (27) Tiraferri, A.; Yip, N. Y.; Phillip, W. A.; Schiffman, J. D.; Elimelech, M., J. Membr. Sci.
49 2011, 367 (1-2), 340.
50
(28) Yip, N. Y.; Tiraferri, A.; Phillip, W. A.; Schiffman, J. D.; Elimelech, M., Environ. Sci.
51
52 Technol. 2010, 44 (10), 3812.
53 (29) Yip, N. Y.; Tiraferri, A.; Phillip, W. A.; Schiffrnan, J. D.; Hoover, L. A.; Kim, Y. C.;
54 Elimelech, M., Environ. Sci. Technol. 2011, 45 (10), 4360.
55 (30) Tiraferri, A.; Kang, Y.; Giannelis, E.; Elimelech, M., Environ. Sci. Technol. 2012,
56 Under review.
57
58 19
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 20 of 29

1
2
3
(31) Tiraferri, A.; Chen, K. L.; Sethi, R.; Elimelech, M., J. Colloid Interface Sci. 2008, 324
4
5 (1-2), 71.
6 (32) Tiraferri, A.; Elimelech, M., J. Membr. Sci. 2012, 389, 499.
7 (33) Staros, J. V.; Wright, R. W.; Swingle, D. M., Anal. Biochem. 1986, 156 (1), 220.
8 (34) Ghosh, A. K.; Jeong, B. H.; Huang, X. F.; Hoek, E. M. V., J. Membr. Sci. 2008, 311 (1-
9 2), 34.
10
11
(35) Wenzel, R. N., Journal of Physical and Colloid Chemistry 1949, 53 (9), 1466.
12 (36) Van Oss, C. J., J. Adhes. Sci. Technol. 2002, 16 (6), 669.
13 (37) van Oss, C. J.; Wu, W.; Docoslis, A.; Giese, R. F., Colloid Surface B 2001, 20 (1), 87.
14 (38) Hurwitz, G.; Guillen, G. R.; Hoek, E. M. V., J. Membr. Sci. 2010, 349 (1-2), 349.
15 (39) Walker, S. L.; Bhattacharjee, S.; Hoek, E. M. V.; Elimelech, M., Langmuir 2002, 18
16
(6), 2193.
17
18 (40) Li, Q. L.; Elimelech, M., Environ. Sci. Technol. 2004, 38 (17), 4683.
19 (41) Butkus, M. A.; Grasso, D., J. Colloid Interface Sci. 1998, 200 (1), 172.
20 (42) Derjaguin, B. V.; Churaev, N. V., Langmuir 1987, 3 (5), 607.
21 (43) Leberman, R.; Soper, A. K., Nature 1995, 378 (6555), 364.
22 (44) Freger, V., Langmuir 2003, 19 (11), 4791.
23
24
(45) Fang, J.; Kelarakis, A.; Estevez, L.; Wang, Y.; Rodriguez, R.; Giannelis, E. P., J.
25 Mater. Chem. 2010, 20 (9), 1651.
26 (46) Mawhinney, D. B.; Glass, J. A.; Yates, J. T., J. Phys. Chem. B 1997, 101 (7), 1202.
27 (47) Wiesner, M. R.; Chellam, S., Environ. Sci. Technol. 1999, 33 (17), 360A.
28 (48) Vrijenhoek, E. M.; Hong, S.; Elimelech, M., J. Membr. Sci. 2001, 188 (1), 115.
29
(49) Tang, C. Y. Y.; Kwon, Y. N.; Leckie, J. O., Desalination 2009, 242 (1-3), 168.
30
31 (50) van Oss, C. J.; Giese, R. F.; Docoslis, A., J. Dispersion Sci. Technol. 2005, 26 (5), 585.
32 (51) Brant, J. A.; Childress, A. E., J. Membr. Sci. 2002, 203 (1-2), 257.
33 (52) Chatterjee, A.; Balaji, T.; Matsunaga, H.; Mizukami, F., Journal of Molecular Graphics
34 & Modelling 2006, 25 (2), 208.
35 (53) He, T.; Ding, H. J.; Peor, N.; Lu, M.; Corley, D. A.; Chen, B.; Ofir, Y.; Gao, Y. L.;
36
37 Yitzchaik, S.; Tour, J. M., J. Am. Chem. Soc. 2008, 130 (5), 1699.
38 (54) Vanoss, C. J.; Chaudhury, M. K.; Good, R. J., Adv. Colloid Interface Sci. 1987, 28 (1),
39 35.
40 (55) Tang, C. Y.; Kwon, Y. N.; Leckie, J. O., J. Membr. Sci. 2009, 326 (2), 526.
41 (56) Lee, S.; Ang, W. S.; Elimelech, M., Desalination 2006, 187 (1-3), 313.
42
43
(57) Adout, A.; Kang, S.; Asatekin, A.; Mayes, A. M.; Elimelech, M., Environ. Sci. Technol.
44 2010, 44 (7), 2406.
45
46
47
48
49
50
51
52
53
54
55
56
57
58 20
59
60
ACS Paragon Plus Environment
Page 21 of 29 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Scheme 1. Schematic of the functionalized nanoparticles and of the protocol to functionalize the thin-
44
45 film composite polyamide forward osmosis membranes used in this study. Polyamide membranes
46
47 possess native carboxylic groups at their surfaces that can be exploited as binding sites for
48 functionalization with tailored nanoparticles. Two different ligands were used to tailor the surface of
49
50 nanoparticles rendering them highly hydrophilic and optimizing their interaction with the membrane
51
surface.
52
53
54
55
56
57
58 21
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 22 of 29

1
2
3
4 Hydrodynamic radius using dynamic light Electrophoretic
5 scattering (nm)* mobility (µm·cm V -1s-1)
6 Deionized water Salt solution Unadjusted pH
7 Bare silica 6.98 ± 0.11 8.15 ± 0.15
8
9 -N(CH3)3+ 8.02 ± 0.16 11.55 ± 0.25 2.48 ± 0.09
10 -NH2/NH3+ 18.99 ± 0.35 20.52 ± 0.45 2.30 ± 0.19
11 *No aggregation observed within 45 minutes
12 0.40

Normalized DTG (% / min)


13 100
Bare silica
C 0.35

Normalized TG (%)
14 99 0.30
15 -N(CH3)3+ 98 0.25
16 0.20
17 A 97
0.15
18 96 0.10
19 95 0.05
20 0.00
21 94
0 100 200 300 400 500 600
22 0.40

Normalized DTG (% / min)


23 100 D 0.35
Normalized TG (%)
24 99 0.30
25
98 0.25
26
0.20
27 97
0.15
28
96 0.10
29 -NH2/NH3+ 95 0.05
30 -N(CH3)3+
31 B 94
0 100 200 300 400 500 600
0.00
32
33 0.40

Normalized DTG (% / min)


34 100
-NH2/NH3+ E 0.35
Normalized TG (%)

35 99 0.30
36 98 0.25
37 0.20
38 97
0.15
39 96 0.10
40 95 0.05
41 0.00
42 94
0 100 200 300 400 500 600
43
Temperature
44
45
46
47 Figure 1. Size, electrophoretic mobility, and thermogravimetric analysis of the functionalized silica
48 nanoparticles. The measured size and electrophoretic mobility of the nanoparticles in deionized water
49 and in an electrolyte solution representative of a typical wastewater effluent (0.45 mM KH2PO4, 9.20 mM
50 NaCl, 0.61 mM MgSO4, 0.5 mM NaHCO3, 0.5 mM CaCl2, and 0.935 mM NH4Cl) are presented in the
51 table. A) and B) show TEM images of silica nanoparticles silanized with –N(CH3)3+-terminated chains
52 and –NH2-terminated chains, respectively. The plots on the right present TGA data for the C) bare silica
53 nanoparticles as well as for the D-E) functionalized nanoparticles. The thermogravimetric plot (line)
54
refers to the left axis and the differential thermogravimetric plot (hollow circles) refers to the right axis.
55
56
Both data sets were normalized by the initial sample mass.
57
58 22
59
60
ACS Paragon Plus Environment
Page 23 of 29 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
A Polyamide O
8
CPS (a.u.)

9
-N(CH3)3+
10 -NH3+/NH2 C
11
12
13 N
14 Si 2s Si 2p
15
16
17 1200 1000 800 600 400 200 0
18 Electron Binding Energy (eV)
19
20 Polyamide
21
22 70 B 70 C 70 D
Area Fraction (%)

23 60 60
-N(CH3)3+ 60
-NH3+/NH2
24 50 50 50
25 40 40 40
26
30 30 30
27
28 20 20 20
29 10 10 10
30 0 0 0
31 O C N O C N Si O C N Si
Element Element Element
32
33
34
35
36
37
38
39
40
41
42 Figure 2. XPS analysis of the surface of the membranes. A) XPS survey scan of control polyamide
43 membranes (black), and of membranes functionalized with silica nanoparticles silanized with –N(CH3)3+-
44
45 terminated chains (red) and –NH2-terminated chains (blue), B-C-D) fractions of oxygen (O, purple),
46
47
carbon (C, green), nitrogen (N, black), and silica (Si, orange) relative to the sum of these elements present
48 at the surface of the three different membranes. The elemental fraction was calculated using software
49
50 CasaXPS from the scans of Figure 2A. The two functionalized membranes show the presence of
51
significant amount of silica at their surface.
52
53
54
55
56
57
58 23
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 24 of 29

1
2
3
4
40
5
Polyamide

Zeta Potential (mV)


6 30
7 20
8 10
9
10 0
11 -10
12 -20
13
-30 A
14
15 4 5 6 7 8 9
16 40
17
-N(CH3)+3
Zeta Potential (mV)
18 30
19 20
20 10
21
0
22
23 -10
24 -20
25 -30 B
26
4 5 6 7 8 9
27
28 40
29 30 -NH+3/NH2
Zeta Potential (mV)

30 20
31
10
32
33 0
34 -10
35 -20
36
37 -30 C
38 4 5 6 7 8 9
39 pH
40
41
42
43
44 Figure 3. Zeta potential of the surface of the membranes as a function of solution pH. A) Zeta potentials
45 of polyamide control membranes, and B-C) zeta potentials of membranes functionalized with silica
46
47 nanoparticles silanized with (red) –N(CH3)3+-terminated chains and (blue) –NH2-terminated chains,
48
49 respectively. Zeta potential values were measured and calculated for at least 4 separately cast and
50 functionalized samples for each membrane type, across a pH range from approximately 4 to 9. The data
51
52 related to different samples were placed in the same plot and represented by different symbols.
53
Measurements were taken at room temperature (23 °C), in solution of 1 mM KCl, and by adjusting the pH
54
55 with appropriate amounts of HCl or KOH.
56
57
58 24
59
60
ACS Paragon Plus Environment
Page 25 of 29 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Figure 4. Surface morphology and roughness of the membranes from SEM and AFM analyses. Surface
44 SEM micrographs of A-B) polyamide control membranes, C-D) membranes functionalized with silica
45
46
nanoparticles silanized with –N(CH3)3+-terminated chains, and E-F) membranes functionalized with silica
47 nanoparticles silanized with –NH2-terminated chains. Panels A, C, and E are low magnification
48 micrographs, while panels B, D, and F are higher magnification surface images. G) representative AFM
49 images of a control polyamide membrane. H) roughness parameters measured by AFM tapping mode
50
51
analysis. Here, RMS is root mean square of roughness, Rmax/10 is maximum roughness divided by a
52 factor of 10, Ra is average roughness, and SAD is percentage surface area difference. Black, red, and blue
53 bars refer to polyamide membranes, and membranes functionalized with –N(CH3)3+- and –NH2-coated
54 nanoparticles, respectively. Roughness values are the average of measurements taken from a total of 12
55
56
random spots on 3 separately cast and functionalized sample surfaces.
57
58 25
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 26 of 29

1
2
3
4
5
A
Contact Angle (degrees)
120 Polyamide
6
7 -N(CH3)3+
8 100
9
Subjected to stress
10 80 Polyamide
11
12 60

on )
.S H
th aO
13

Cl

.
40
Na

pH Cl)
Ba 2 (N
14

(H
M
15

1
2
6

20
0.

pH
16
17
18 0
19
-N(CH3)3+
B
20
Contact Angle (degrees)

21 120 Polyamide
22 -NH2/NH3+
23 100
24 Subjected to stress
25 80
26 -NH2/NH3+
. S H)
O

27 60
12 l)
2 l
pH NaC

.
pH (HC

Ba (Na

on

28
M

29
th

40
6
0.

30
31
32
20
33
34 0
35
36
37
38
39 Figure 5. Contact angles of deionized water on the surface of the membranes for A) membranes
40
41 functionalized with silica nanoparticles silanized with –N(CH3)3+-terminated chains (red), and B)
42 membranes functionalized with silica nanoparticles silanized with –NH2-terminated chains (blue). The
43
44 contact angle of DI water on control polyamide membranes is shown in both plots as a patterned black
45
bar. The plots show values of the membranes as functionalized (solid bars), and after the surface was
46
47 subjected to stress (hollow bars), as briefly labeled in the graphs on each bar and as described in the
48
49 discussion. Values are averages of at least 8 random spots from each sample. Measurements were
50 carried out at room temperature (23 °C), without addition of ionic strength, and at unadjusted pH. When
51
52 contact angles were too low to be accurately measured, a value of 10 degrees was assumed for the
53
54 calculations. Representative pictures of DI water droplets are included on the right for illustration
55 purposes.
56
57
58 26
59
60
ACS Paragon Plus Environment
Page 27 of 29 ACS Applied Materials & Interfaces

1
2
3
4
5 -∆G13 (Wettability)
6

Specific Energy (mJ/m2)


7 Polyamide
8
120 -N(CH3)3+
9 -NH2/NH3+
10
80
11
12
40
13 0
14
15 -40
16
17
18
-80 A
19
20 -∆G131 (Hydrophilicity)
21
22
23 60
Surface Energy (mJ/m2)

24
25 50
B
26
27 40
28
29 30
30
31
32
20
33
34 10
35
36 0
37
38
39
40
41
42
43 Figure 6. Wettability, hydrophilicity, and surface energy of the surface of the membranes. A) wettability
44 with DI water, -ΔGML, and of hydrophilicity, ΔGTOTMLM, and B) calculated values of surface energy, γTOT.
45
46 Data for polyamide control membranes are presented as black patterned bars. Values for membranes
47
functionalized with silica nanoparticles silanized with –N(CH3)3+-terminated chains or with –NH2-
48
49 terminated chains are shown as red and blue bars, respectively. The surface energy parameters were
50
51 calculated from average contact angles measured with DI water, glycerol, and diiodomethane at room
52 temperature (23 °C), without addition of ionic strength, and at unadjusted pH. At least 25 contact angles
53
54 on at least 3 separately cast and functionalized samples were measured for each liquid and for each
55
56 membrane type.
57
58 27
59
60
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 28 of 29

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 7. Representative AFM retraction curves for foulant-membrane interaction using a A) BSA-
42
43 fouled tip, and B) alginate-fouled tip. Data for control polyamide and for membranes functionalized with
44
45 –N(CH3)3+-terminated nanoparticles are shown as black squares and red circles, respectively. The
46 average, minimum, and maximum values of the minimum energy wells measured for 125 separate
47
48 retracting curves are reported for each foulant. The ‘No’ label stands for measurements where no
49
adhesion force was observed. The test solution for the measurements is synthetic wastewater as described
50
51 in the experimental section. Measurements were carried out at room temperature (23 °C).
52
53
54
55
56
57
58 28
59
60
ACS Paragon Plus Environment
Page 29 of 29 ACS Applied Materials & Interfaces

1
2
3
Table 1. Summary of the contact angle and surface energy data of the different membranes
4
5 analyzed in this study. Average contact angles of the water, glycerol, and diiodomethane are
6
7 reported (degrees), along with the different components of the surface energy of the membrane
8
9 surface, expressed in mJ/m2
10
11
12
13
14
15
Membrane θwat θgly θdiod SAD LW ΔGMLM
16 γ γ+ γ- γAB γTOT -ΔGML
17 (%) (TOT)
18
19 Polyamide 105 76.5 27.2 23.0 37.7 0.06 0.03 0.09 37.8 57.9 -97.9
20
21
22 -N(CH3)3+ <10 17.6 18.3 19.3 41.0 1.10 38.3 13.0 54.0 133 +12.3
23
24
25 -NH3+/NH2 19.9 23.7 25.7 9.9 42.1 1.2 39.4 13.9 55.9 135 +12.7
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 29
59
60
ACS Paragon Plus Environment

You might also like