PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE SI08 KzlProj/Acc8Pres/CIRC/Dateoue.indd COMBINED FOULING OF PRESSURE-DRIVEN MEMBRANES TREATING FEED WATERS OF COMPLEX COMPOSITION By Fulin Wang A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Environmental Engineering 2008 ABSTRACT COMBINED FOULING OF PRESSURE-DRIVEN MEMBRANES TREATING FEED WATERS OF COMPLEX COMPOSITION By Fulin Wang Pressure-driven membrane filtration processes can serve as a reliable barrier for microorganisms, organics, precursors of disinfection by-products (DBPs), colloids, and other contaminants. However, membrane fouling impedes a wider application of membrane filtration technology. Because typical feed waters have complex composition and the pretreatment often can not guarantee a complete removal of foulants, different fouling mechanisms act simultaneously resulting in combined fouling. Although each of the individual fouling mechanisms has been extensively studied, there have been very few studies examining the effect of combined fouling on membrane performance. The focus of this work is on identifying contributions of individual fouling mechanisms in the overall fouling process and on evaluating the synergistic effects stemming from interactions between individual mechanisms. In Chapter 2, concentration of rejected salt at the membrane surface when colloidal particles were deposited on the membrane is determined experimentally based on measured salt permeability constant. This approach allowed for a clear identification of individual contributions of concentration polarization and colloidal fouling to the permeate flux decline. In Chapter 3, a method based on a simple linear regression n fitting is proposed and used to determine the type, the chronological sequence, and the relative importance of individual fouling mechanisms in the experiments on the dead-end membrane filtration of colloidal suspensions. For all membranes, flux decline was consistent with one or more pore blocking mechanisms during the earlier stages and with the cake filtration mechanism during the later stages of filtration. In Chapter 4, a model is proposed for predicting the permeability of porous media consisting of fibers and spherical particles. The model combines Kozeny-Carman equation and Ethier’s theoretical model to predict the permeability of the mixed media based on the permeability of the porous bed of particles alone and on the volume fraction of the fibers. The model prediction was found to be in accordance with the experimental observations for the fiber mass fractions of less than 50%. Chapters 5 and 6 examine the effect of combined colloidal fouling and gypsum scaling on the performance of RO membranes. In Chapter 5, the effect of silica colloids on gypsum scaling is studied in a batch system. The results showed that silica colloids retarded the induction of gypsum crystallization and decreased the nucleation rate. It was also found that gypsum crystallization rate increased in the presence of silica colloids. Chapter 6 presents the first study on the combined colloidal fouling and gypsum scaling in RO membrane systems. A significant synergistic effect between gypsum scaling and colloidal fouling was observed. To my dearest parents, my brother, and my sister. iv ACKNOWLEDGEMENTS I would like to express my deepest gratitude and indebtedness to my advisor and dissertation committee chairman, Dr. Volodymyr Tarabara, for his continuous mentoring, support, and guidance during this work. I would also like to thank him for his generous encouragement, infectious enthusiasm and constructive criticisms, without which the work described in this dissertation would not have been possible. I would also like to extend my great appreciation to the other dissertation committee members: Dr. Phanikumar S. Mantha, Dr. Merlin L. Bruening, and Dr. Manoochehr Koochesfahani for their valuable comments and suggestions on this work. I would like to thank Dr. Seokjong Byun, Dr. Jeonghwan Kim, Julian Taurozzi, and Wenqian Shan, Alla Alpatova, Adam Rogensues, Elodie Pasco, Tan Zhao, for many fi'uitful discussions and for their kind help in sharing lab equipment. I would like to acknowledge Ms. Lori Hasse and Mr. Yang-lyang Pan for the administrative and research supports. I also would like to thank Ms. Ewa Danielewicz, Dr. Alicia Pastor, and Ms. Carol S. Flegler for their constant assistance with microscopy sample preparation. I would like to thank the Department of Civil and Environmental Engineering, Michigan State University, and the National Water Research Institute for the financial support of this work. Finally I would like to thank my parents, my brother, and my sister for their enduring love, support, sacrifice, and motivation over the years. vi TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. xi LIST OF FIGURES .......................................................................................................... xii ABBREVIATIONS .......................................................................................................... xv CHAPTER ONE COMBINED FOULING ON THE PERFORMANCE OF PRESSURE-DRIVEN MEMBRANES TREATING WATERS OF COMPLEX COMPOSITION: CURRENT UNDERSTANDING AND FUTURE RESEARCH NEEDS 1 .1. Introduction .................................................................................................................. 1 1.1.1. Overview of membrane filtration processes ......................................................... 1 1.1.2. Combined fouling in practice ................................................................................ 3 1.2. Current research on the effect of combined fouling on membrane performance ........ 6 1.2.1. Colloidal fouling and salt concentration polarization in salt-rejecting membrane systems.. ......................................................................................................................... 6 1.2.1.1. Colloidal fouling ............................................................................................ 6 1.2.1.2. The structure of colloidal deposit layers ........................................................ 9 1.2.1.3. Coupled effect of colloidal deposition and salt concentration polarization. 10 1.2.2. Combined colloidal and organic fouling ............................................................. 12 1.2.3. Colloidal fouling and scaling .............................................................................. 14 1.2.4. Other combined fouling ...................................................................................... 16 1.3. The future research need ............................................................................................ 17 Reference .......................................................................................................................... 18 CHAPTER TWO COUPLED EFFECTS OF COLLOIDAL DEPOSITION AND SALT CONCENTRATION POLARIZATION ON REVERSE OSMOSIS MEMBRANE PERFORMANCE ABSTRACT ...................................................................................................................... 27 2. 1. INTRODUCTION ..................................................................................................... 28 2.2. BACKGROUND ....................................................................................................... 32 2.3. EXPERIMENTAL METHODS ................................................................................. 37 2.3.1. Experimental apparatus ....................................................................................... 37 2.3.2. Reagents and chemical solutions ........................................................................ 40 2.3.3. Silica colloids and R0 membranes ..................................................................... 41 2.3.4. Experimental protocol ......................................................................................... 42 2.3.4.1. Membrane compaction ................................................................................. 42 vii 2.3.4.2. Pure water permeate flux and measurement of membrane hydraulic resistance ................................................................................................................... 43 2.3.4.3. Membrane conditioning ............................................................................... 43 2.3.4.4. Membrane fouling experiments ................................................................... 44 2.4. RESULTS AND DISCUSSION ................................................................................ 45 2.4.1. Characteristics of colloidal particles and the membrane .................................... 45 2.4.2. Changes in membrane resistance during conditioning. Correction factor .......... 46 2.4.3. Measurement of the membrane salt permeability constant ................................. 49 2. 4. 4. Permeate flux, NaCl rejection, and structure of the deposited layer at different ionic strengths ................................................................................ 51 2. 4. 5. Hypothesis of the concentrated flowing layer of silica partibles and its depolarizing effect on NaCl transport ........................................................................... 59 2.4.6. Evolution of relative contributions of osmotic pressure and resistance of the colloidal deposit to the permeate flux decline .............................................................. 60 2.4.7. Effect of the feed channel spacers on NaCl rejection and permeate flux ........... 61 2.5. CONCLUSIONS ........................................................................................................ 65 ACKNOWLEDGEMENTS .............................................................................................. 66 REFERENCE .................................................................................................................... 68 82. SUPPLEMENTARY DATA ..................................................................................... 72 8.2.1. Determination of the salt permeability ............................................................... 72 8.2.2. Imaging the fouling layer using transmission electron microscopy (TEM) ....... 73 CHAPTER THREE ON THE IDENTIFICATION OF PORE BLOCKING MECHANISMS DURING EARLY STAGES OF MEMBRANE FOULING BY COLLOIDS ABSTRACT ...................................................................................................................... 75 3. 1 . INTRODUCTION ..................................................................................................... 76 3.2. Materials and methods ............................................................................................... 81 3.2.1. Membranes and colloids ..................................................................................... 81 3.2.2. Filtration experiments ......................................................................................... 82 3.2.3. Scanning electron microscopy (SEM) analysis .................................................. 84 3.2.4. Data fitting procedure ......................................................................................... 84 3.2.5. Determination of the effective porosity of the colloidal deposit ........................ 85 3.3. Results and discussion ............................................................................................... 86 3.3.1. Flux performance of different membranes ......................................................... 86 3.3.2. Blocking mechanisms for UF membranes .......................................................... 89 3.3.3. Blocking mechanisms for salt—rejecting membranes .......................................... 91 3.4. Conclusions ................................................................................................................ 96 Acknowledgements ........................................................................................................... 97 Reference .......................................................................................................................... 98 S3. SUPPLEMENTARY DATA .................................................................................... 101 viii CHAPTER FOUR MODELING THE PERMEABILITY OF FIBER-FILLED POROUS MEDIA USING KOZENY-CARMAN-ETHIER EQUATION 4. 1. INTRODUCTION ................................................................................................... 108 4.2. Theoretical background ........................................................................................... 110 4.2.1. The hydraulic resistance of a deposit layer in membrane systems ................... 110 4.2.2. The hydraulic resistance of porous media of particles ...................................... 112 4.2.3. The permeability of highly polydisperse fibrous media ................................... 112 4.3. Experimental method ............................................................................................... 113 4.3.1. Experimental apparatus ..................................................................................... 1 13 4.3.2. Reagents and chemicals .................................................................................... 116 4.3.3. Ultrafiltration (UF) membrane, Silica colloids, and carbon fibers .................... 116 4.3.4. Scanning electron microscopy (SEM) analysis ................................................ 117 4.3.5. Experimental protocol ....................................................................................... 117 4.3.5.1. Preparation of the feed suspensions ........................................................... 117 4.3.5.2. Membrane compaction ............................................................................... 117 4.3.5.3. Measurement of membrane hydraulic resistance using pure water ........... 118 4.3.5.4. Membrane filtration experiments ............................................................... 118 4.4. Results and discussion ............................................................................................. 119 4.4.1. Characterization of the silica colloids and carbon fibers .................................. 119 4.4.2. Permeate flux and hydraulic resistance of the deposit layer ............................. 122 4.4.3. SEM images of the deposit layers ..................................................................... 124 4.4.4. Modeling the hydraulic resistance of the deposit layers consisting of silica colloids and fibers ....................................................................................................... 124 4.4.4.1. Theoretical development ............................................................................ 124 4.4.4.2. The effective porosity of the porous bed of silica colloids ........................ 128 4.4.4.3. Prediction of the hydraulic resistance of the deposit layer consisting of silica colloids and carbon fibers ....................................................................................... 131 4.5. Conclusions .............................................................................................................. 134 Acknowledgments ........................................................................................................... 135 Nomenclature .................................................................................................................. 136 Reference ........................................................................................................................ 138 CHAPTER FIVE CRYSTALLIZATION OF GYPSUM IN THE PRESENCE OF COLLOIDAL SILICA. I: IN A STIRRED BATCH SYSTEM ABSTRACT .................................................................................................................... 141 5. 1 . Introduction .............................................................................................................. 143 5.2. Background .............................................................................................................. 146 5.3. Material and methods ............................................................................................... 151 5.3.1. Experimental apparatus ..................................................................................... 151 5.3.2. Regents, silica colloids, and feed suspensions .................................................. 153 ix 5.3.3. Characterization of the gypsum crystals ........................................................... 155 5.4. Results and discussion ............................................................................................. 156 5.4.1. Determination of induction time ....................................................................... 156 5.4.2. Effect of silica colloids on the induction of gypsum crystallization ................. 159 5.4.3. Effect of silica colloids on the surface energy and nucleation rate of gypsum crystallization .............................................................................................................. 163 5.4.4. Effect of silica colloids on the gypsum crystallization rate .............................. 165 5.4.5. Effect of silica colloids on crystal size distribution .......................................... 166 5.4.6. Effect of silica colloids on the morphology of gypsum crystals ....................... 166 5.4.6.1. Length-to-width ratio ................................................................................. 168 5.4.6.2. Implications of epitaxial growth of gypsum for crystal morphology ........ 169 5.5. Conclusions .............................................................................................................. 171 Acknowledgments ........................................................................................................... 172 Nomenclature .................................................................................................................. 173 Reference ........................................................................................................................ 174 SS. SUPPLEMENTARY DATA .................................................................................... 179 CHAPTER SIX CRYSTALLIZATION OF GYPSUM IN THE PRESENCE OF COLLOIDAL SILICA. 11: IN A REVERSE OSMOSIS MEMBRANE SYSTEM Abstract ........................................................................................................................... 183 6.1 . Introduction .............................................................................................................. 184 6.2. Experimental ............................................................................................................ 187 6.2.1. Feed suspensions ............................................................................................... 187 6.2.2. Experimental apparatus ..................................................................................... 189 6.2.3. Reagents, silica particles and R0 membranes .................................................. 190 6.2.4. Microscopical analysis ...................................................................................... 190 6.2.5. Protocols for membrane fouling experiments with less saturated suspensions 191 6.2.6. Study the effect of silica particles on gypsum crystallization with supersaturated suspensions ................................................................................................................. 1 92 6.3. Results and discussion ............................................................................................. 193 6.3.1. Combined colloidal fouling and gypsum scaling in less saturated suspensions 193 6.3.1.1. Effect on RO membrane performance ....................................................... 193 6.3.1.2. Effect of colloidal fouling on gypsum scaling ........................................... 198 6.3.1.3. Effect of gypsum scaling on colloidal fouling ........................................... 200 6.3.1.4. Microscopical analysis of the fouling layer on membrane surface ............ 200 6.3.2. Combined colloidal fouling and gypsum scaling in supersaturated suspensions ..................................................................................................................................... 203 6.3.2.1. Effect of silica particles on gypsum crystallization ................................... 203 6.3.2.2. Effect on RO membrane performance ....................................................... 207 6.3.2.3. Microscopical analysis of the fouling layer on membrane surface ............ 207 6.4. Conclusions .............................................................................................................. 210 Acknowledgments ........................................................................................................... 21 1 Reference ........................................................................................................................ 212 LIST OF TABLES Table 1.1. Examples for combined fouling occurring in practice ....................................... 5 Table 1.2. Published studies of combined fouling in membrane systems .......................... 7 Table 2.1. R0 experimental protocol ................................................................................ 48 Table 2.2. Measurement of the membrane salt permeability constant ............................. 50 Table 82.1. Measurement of the membrane salt permeability constant ........................... 72 Table 3.1. Membranes used in this study .......................................................................... 83 Table 4.1. Rm in different experiments .......................................................................... 123 Table 5.1. Composition of model solutions .................................................................... 154 Table 5.2. Estimated critical size of nuclei in the absence and presence of silica colloids ................................................................................................................................. 162 xi LIST OF FIGURES Figure 2.1. Schematic of the experimental apparatus ....................................................... 39 Figure 2.2. Flux change during conditioning, at a transmembrane pressure difference of 300 psi (2.068 MPa) and an ionic strength of 10'2 M. .............................................. 49 Figure 2.3. (a) Normalized permeate flux at three ionic strengths;, (b) observed salt rejection; (c) salt concentration at membrane surface; ((1) mass of deposited particles; (e) effective porosity, and (f) hydraulic resistance of deposited colloidal layer ....... 52 Figure 2.4. Effect of introduction of silica particles on initial observed salt rejection at 10'2 M ........................................................................................................................ 57 Figure 2.5 Energy consumption analysis for the system: ................................................. 62 Figure 2.6. Effect of the feed channel Spacers on membrane performance ...................... 64 Fig. 82.1. TEM imaging of the deposit layer formed at 0.1 M ......................................... 73 Figure 3.1. Schematic illustration of the four fouling mechanisms: (a) complete blocking, (b) standard blocking, (c) intermediate blocking, and ((1) cake filtration. ................ 77 Figure 3.2. Normalized instantaneous permeate flux (a) and revere cumulative flux (b) as a function of permeate volume for different membranes .......................................... 88 Figure 3.3. The distributions and contributions of the blocking mechanisms in terms of permeate volume (a) and percentage of flux decline (b). ......................................... 90 Figure 3.4. SEM images of the filtration surfaces of membranes. (a) BW30-365; (b) NF90; (c) UF-30kDa; (d) UF-lOOkDa; and (e) UF-300kDa. ................................... 93 Figure 3.5. Effective porosity of the layer of silica colloids deposited onto the surface of BW30-365 and NF90 membranes. ........................................................................... 95 Fig. S3. 1. Application of blocking laws to the filtration data of UFlOOkDa membrane: (a) 1/ Q versus t based on Eq. (3.7) for intermediate blocking; (b) t/ V versus I based on Eq. (3.2) for standard blocking; and (c) t/V versus V based on Eq. (3.4) for cake filtration. ......................................................................................................... 101 Fig. 832. Application of blocking laws to the R0 and NF membrane filtration data: t/ V versus V based on Eq. (3.4) for cake filtration. ..................................................... 102 xii Fig. 83.3. Application of blocking laws to the R0 and NF membrane filtration data: (a) Q versus V based on Eq. (3.6) for complete blocking, and (b) 1/ Q versus I based on Eq. (3.7) for intermediate blocking. ................................................................... 103 Figure 4.1. Idealized cross-section view of the porous matrix ....................................... 115 Figure 4.2. SEM images for the silica colloids (a) and carbon fibers (b) ....................... 120 Figure 4.3. Permeate flux (a) and hydraulic resistance of the deposit layer (b). ............ 121 Figure 4.4. SEM of the deposition layers (continued on next page) ............................... 125 Figure 4.4. SEM of the deposition layers. ...................................................................... 126 Figure 4.5. Porosity of the silica deposit layer as a fiinction of permeate volume ......... 130 Figure 4.6. Comparison between model prediction and experimental measurement for the hydraulic resistance of the deposition layer ............................................................ 133 Figure 5.1. Schematic of the experimental batch system ................................................ 152 Figure 5.2. Comparison of different techniques for measuring gypsum induction time 157 Figure 5.3. The transmittance (a) and conductivity (b) of the feed solutions as a function of time ..................................................................................................................... 160 Figure 5.4. Effect of silica colloids on (a) induction time, (b) nucleation rate, and (c) crystallization rate ................................................................................................... 164 Figure 5.5. Size distribution of gypsum crystals formed in the absence and in the presence of 50 mg/L silica colloids. ...................................................................................... 167 Figure 5.6. SEM image of gypsum crystals formed in the absence (a-d) and in the presence (e-h) of 50 mg/L silica colloids ................................................................ 170 Figure SS. 1. Transmittance as a function of time at three wavelengths ......................... 179 Figure 85.2. Effect of particles on transmittance and conductivity ................................ 180 Figure 85.3. The plots for calculating the surface energy of the crystals ....................... 181 Figure 6.1. Comparison of membrane flux decline in the RO membrane system during gypsum scaling (indicated as “CaC12+ Na2804 only”), colloidal fouling (“CaC12+ NaCl + particles”), and combined fouling experiments (“CaC12+ NaZSO4 + particles”). ............................................................................................................... 194 xiii Figure 6.2. Permeate conductivity of the RO membrane during gypsum scaling, colloidal fouling, and combined fouling experiments, respectively. ..................................... 196 Figure 6.3. Mass of silica particles deposited in the absence and presence of gypsum scaling ..................................................................................................................... 199 Figure 6.4. Gypsum crystals in the absence of silica particles ....................................... 201 Figure 6.5. Gypsum crystals in the presence of silica particles at different magnifications ................................................................................................................................. 201 Figure 6.6. X-ray energy dispersive spectroscopy (EDS) analysis of the fouling layer consisting of gypsum crystals and silica particles .................................................. 202 Figure 6.7. Effect of silica particles on gypsum crystallization at different operating conditions ................................................................................................................ 205 Figure 6.8. Normalized permeate flux at three experimental conditions ........................ 206 Figure 6.9. The crystals formed in the absence of silica particles. ................................. 208 Figure 6.10. The gypsum crystals formed in the presence of silica particles. ................ 209 xiv AF M DAF F DBPs DLVO EDS EPS MF MWCO NF NOM PA PAN PES PTFE PVDF R0 SDI TEM TDS TOC TMP SEM W254 ABBREVIATIONS Atomic force microscopy Dissolved air flotation and filtration Disinfection by-products Deijaguin-Landau-Verwey-Overbeek Energy dispersive spectroscopy Extracellular polymeric substances Microfiltration Molecular weight cutoff Nanofiltration Natural organic matter Polyamide Polyacrylonitrile Polyethersulfone Polytetrafluoroethylene Polyvinylidenefluoride Reverse osmosis Silt density index Transmission electron microscopy Total Dissolved Solids Total organic matter Transmembrane pressure difference Scanning electron microscopy Ultrafiltration Ultraviolet absorbance measured at 254 nm XV CHAPTER ONE COMBINED FOULING ON THE PERFORMANCE OF PRESSURE- DRIVEN MEMBRANES TREATING WATERS OF COMPLEX COMPOSITION: CURRENT UNDERSTANDING AND FUTURE RESEARCH NEEDS 1.1. Introduction Our growing demand for drinking water has led many water utilities to use alternative water sources, such as brackish surface water and seawater, with elevated levels of contaminants and salts [1-3]. Meanwhile, the increasingly stringent water standards, such as Disinfectants/Disinfectants Byproducts Rule (D/DBP) and Long Term 2 Enhanced Surface Water Treatment Rule (LT2ESWTR), require a higher removal of contaminates [4]. Membrane filtration processes are continuously gaining popularity because they can serve as a reliable means for removing microorganisms, organics, precursors of disinfection by-products (DBPs), colloids, and other suspended matter of concern [4-6]. 1.1.1. Overview of membrane filtration processes Pressure-driven membrane filtration processes, including microfiltration (MF), ultrafiltration (U F), nanofiltration (NF), and reverse osmosis (R0), are being increasingly used by water and wastewater treatment utilities [4, 6]. Low pressure membranes, MF and UF are primarily used for removing turbidity, pathogens, and particles from fresh waters [7]. They are used as stand-alone processes [8, 9] or as pretreatment for high pressure salt—rejecting NP and R0 membranes [10, 11]. NF or R0 membranes are usually used for softening, desalination, and removal of natural organic matter (NOM) and precursors of DBPs [7]. The major obstacles to a wider application of membrane filtration technology are membrane fouling and disposal of concentrate. Membrane fouling occurs when the rejected contaminants accumulate within the membrane or on the membrane surface, resulting in a decrease in permeate flux. Membrane fouling shortens membrane life, decreases both permeate quantity and quality, and increases capital and operational costs [6, 12]. Depending on the properties of the contaminants being rejected by a membrane, membrane fouling can be roughly categorized into four types [7]: (l) Particulate fouling by particles in colloidal range (colloidal fouling) or larger; (2) Organic fouling by NOM and other organic molecules; (3) Biological fouling from the accumulation or growth of microorganisms; and (4) Scaling resulting from sparingly soluble salts such as carbonate and sulfate. Understanding membrane fouling is crucial to developing low-fouling membranes, controlling membrane fouling, and ultimately reducing operating costs. However, our knowledge of membrane fouling is limited. For example, theoretical predictions of permeate flux underestimate the real membrane performance, the phenomenon is known as the “flux paradox” [13]. An accurate estimation of permeate flux still remains elusive F_' A—‘ D— in part due to the failure of predicting the fouling layer structure [14]. This is mainly due to the fact that, in practice, fouling is caused by several foulants and involving different fouling mechanisms simultaneously. 1.1.2. Combined fouling in practice It was reported that treated wastewater presently accounts for 5%, brackish water for 22% and seawater for 58% of the water produced by desalination technologies, among which the membrane filtration process is predominating [15]. Since the first high-flux, anisotropic reverse osmosis membranes were invented in the early 1960s [16], NF and R0 membranes have played an increasingly important role for desalination. Brackish water is defined as water containing total dissolved solids (TDS) levels that are higher than potable water, but lower than seawater (in the range of 1,000 to 25,000 mg/l TDS) [l7]. Waters found in coastal areas, aquifers, surface waters, and groundwater can be brackish [2, l7]. Brackish water usually has a complex composition. For example, the Colorado River, the primary water source for the American Southwest and the feed water for many membrane treatment plants, has a complex composition [18] and includes high amounts of organic matter, colloids, and salts. For seawater, in addition to a high concentration of salts, a substantial amount of inorganic particles of a wide range of sizes is present [19, 20]. The particulate matter includes elemental sulfur, silica, and ferric and aluminum hydroxides [19, 21-24]. Although feed waters for NF/RO membrane systems usually are pretreated and thus the colloidal fraction is largely removed, the performance of the pretreatment systems varies seasonally [25-27] and cannot guarantee a complete removal of components of small size [19—21, 24, 28]. Boerlage and co-workers [26] evaluated colloidal fouling in two RO pilot desalination plants. In these two plants, the feed waters were pretreated by a slow sand filtration and a UF membrane system, respectively, before entering the RO membrane system. It was found that both of the pretreatment methods were effective in removing particles larger than 0.45 pm, but not effective in removing the smaller particles. Based on ca. 2 years of monitoring in a seawater R0 plant, Chua et al. [19] reported that although both a conventional pretreatment system and an ultrafiltration (UF) membrane system were used to remove the suspended particles [19], the silt density index (SDI) of the feed water for the RO membrane system remained above 2.5 and a substantial quantity of suspended particles were not removed. Examples of combined fouling observed in practice are summarized in Table 1.1. Table 1.1. Examples for combined fouling occurring in practice Feed water Location Membrane Pretreatment Foulants Ref. Oise river, NF - Organics, [29] Paris, France colloids, scale United R0 Ozonation, Biological [30] _ Kingdom dissolved air matter, iron, Bmk‘Sh flotation, silica colloids surface sofiening, media water filtration, and cartridge filter Colorado RO MF Colloids and [31] River, USA scale Zarzis and R0 Disinfection, sand Organic matter, [32, Djerba Island, filtration, silica, iron, and 33] Brackish Tunisia oxidation, and MF scale groundwater Minjur aquifer, R0 Media filter, and Iron, silica [34] Saudi Arabia cartridge filter colloids, scale Seawater Taiwan Strait, R0 Media filtration, Organic matter, [3 5] mixed with Taiwan and cartridge filter scales, and EPS groundwater Umm Lujj, NF Media filtration, Organic matter, [36] Saudi Arabia and cartridge filter iron, scale, chromium and Seawater fungus Eilat, Israel R0 Coagulation, UF Iron, chromium, [3 7] (or sand and organic and cartridge filter) biological matter Sydney, MF Biological Colloids, organic [38] Australia treatment matter, and scale , _ Greensboro, RO/N F MF/UF Organic, [39] MunIcrpal NC, USA inorganic, and wastewater biological matter California, R0 MF/UF NOM, colloidal [40] USA silicates, and scale 1.2. Current research on the effect of combined fouling on membrane performance The studies published to date on studying the effect of combined fouling so far are listed in Table 1.2. Surprisingly, although the fouling mechanisms have been extensively examined individually, very few studies have been conducted on the combined fouling. 1.2.1. Colloidal fouling and salt concentration polarization in salt-rejecting membrane systems 1.2.1.1. Colloidal fouling During membrane filtration, colloids can adhere to the inner pores of porous membranes and/or form a cake on membrane surface. Colloidal fouling increases the hydraulic resistance to the permeate flow and thus decreases permeate flux. The two dominating factors determining the effect of colloidal fouling on membrane performance are 1) the amount of colloids on the membrane surface, and 2) structure of the deposited layer (cake). Colloidal deposition is influenced by hydrodynamics (crossflow rate, transmembrane pressure or initial permeate flux), solution chemistry (ionic strength and composition, and pH), membrane properties (surface charge, roughness, and hydrophilicity/hydrophobicity), and colloid properties (particle size and charge). Chellam and Wiesner [49] studied particle transport and deposition in laminar crossflow Table 1.2. Published studies of combined fouling in membrane systems Combined fouling Membrane References Colloidal and organic fouling MF, UP, and NF [41, 42][43][44-46] Colloidal fouling and scaling R0 [47] Organic fouling and scaling NF [48]* *Brackish water and synthetic water were used. In other studies, only synthetic water was used. membrane filtrations and found that the mechanism controlling lateral migration depended on inertial lifi, gravity, and permeation drag forces. Belfort et al. [13] extensively reviewed back-transport of particles away from the membrane surface due to Brownian diffusion, shear-induced diffusion and inertial lifi. Higher crossflow rate and lower transmembrane pressure (or initial permeate flux) help to alleviate colloidal fouling [50, 51]. A faster crossflow produces greater shear at the membrane surface, and thus increases the back diffusion of particles into the bulk suspension [52]. A lower transmembrane pressure results in a slower initial permeate flow, which brings a small amount of colloids to the membrane surface. This phenomenon leads to the concept of “critical flux”, at which no colloidal fouling occurs due to the fact that the back diffusion and convection of the colloids balance their deposition onto the membrane surface ([53] and the references therein). Colloidal deposition is usually more severe when the membrane has a rougher surface [54, 55]. Boussu et al. [56] has demonstrated that roughness played an important role only when the colloids were relatively small. 0n the contrary, Van der Bruggen et al. [5 7] have shown no correlation between surface roughness and colloidal deposition rate. Surface charge and hydrophilicity/hydrophobicity have important effects on colloidal deposition as well. Generally speaking, hydrophilic membranes with higher surface charge are preferred to minimize the colloidal adhesion to the membrane surface [16, 56, 58, 59]. Faibish et al. [51] demonstrated that small particles resulted in a higher flux decline than large particles at the same particle loading because smaller size resulted in a higher hydraulic resistance and thus a lower permeability. Van der Bruggen et a1. [46] observed the same trend in MF membrane filtration, which was due to the fact that the smal 1 particles blocked the pores and reduced permeate flux rate. However, in NF membrane filtration from the same study, large particles fouled membrane more severely, which was attributed to the slower back diffusion of the large particles from the membrane surface to the bulk solution. 1.2. l -2. The structure of colloidal deposit layers The structure of colloidal deposition layer is known to be a complex firnction of operational variables such as hydrodynamics, colloid size and charge, and solution chemistry [60-65]. Solution chemistry-mediated interactions between particles within the deposited layer determine the structure and, therefore, specific hydraulic resistance of the deposit [66]. Ionic strength influences the surface charge and thus the structure of the Colloidal deposit layer as well [60]. At a high ionic strength, the repulsive forces between the particles are reduced, which is attributed to the compression of the ionic double layer surrounding particles, as predicted by the classic Derjaguin-Landau-Verwey-Overbeek (DLVQ) theory. As a result, the porosity of the deposit layer is usually lower and COHOi dal fouling is more severe at higher ionic strengths. This has been confirmed by numerous studies (e.g. [50, 64, 67, 68]). Similar to the effects of ionic strength, pH influences colloidal fouling by changing the surface charge of the particles and membranes. The effect of pH is strongly dependent on the isoelectric point of the pafilcles and membranes. The further deviation from the isoelectric point (resulted from either increasing pH above it or decreasing pH below it) usually increases the absolute value of the surface charge of the particles and membranes [50, 69]. Qin et a1. [70] demonstrated that a lower pH led to less fouling in a RO membrane system for reclamation of spent rinse water from metal plating. F aibish at el. [67] observed that pH influenced the initial permeate flux but not the stable permeate flux. The higher initial permeate flux at lower pH was contributed to the apparent viscosity of the permeating solution near the surface. Theoretical and experimental studies indicate that colloidal cakes can be stratified and can contain layers with dense and loose particle packing [61, 63, 64, 71]. It has been also reported that the porosity of the colloidal fouling layer during the cross-flow filtration decreased with filtration time [72]. Although the cake structure was not measured Continuously from the cross-flow membrane unit, the results suggest that the structure of the Fouling layer should be time-dependent. It is important to note that, in practice, these effects are expected to act simultaneously. For instance, change in pH or ionic strength affects both membrane and particle charges, WhiCh Changes the interaction forces between membranes and colloids and also among the colloids themselves [56]. 1-2-1—3 - Coupled effect of colloidal deposition and salt concentration polarization When a solute is rejected by a membrane, the concentration of the solute at the membrane surface will be higher than solute concentration in the bulk of the feed solution. This 10 leads to the formation of a gradient of solute concentration toward the membrane surface This phenomenon is known as concentration polarization. When colloidal fouling is present, the fouling layer formed on the membrane surface hinders the diffusion of solutes in the layer [73-75], which increases the solute concentration at the membrane surface. The hindrance effect is strongly dependent on the structure of the layer [74], which, as discussed above, is a complex function of flow hydrodynamics, colloid size and charge, and solution chemistry. Rejection of NaCl by RO membranes was shown to decrease in the presence of membrane fouling by iron hydroxide [76, 77] and by silica colloids [76, 78-80]. In a Study of effects of nanofiltration membrane fouling on rejection, Shafer et al. [81] 1' eported that rejection of cations and negatively charged low molecular weight acids depended on the charge of ferric chloride precipitate deposited on the membrane surface. Hoek et al. [78, 79] have proposed a model that described how hindrance of the back- diffusion of salt by a stagnant colloidal cake results in cake-enhanced concentration POIariZation, an increase in osmotic pressure, and, consequently, lower fluxes and salt rejeCtiOns. Results of the model, which assumed the porosity of the colloidal silica layer to be constant, agreed well with experimental data on permeate flux and salt rejection. A modified cake-enhanced concentration polarization model was later proposed by Ng and Elimelech [82], wherein the colloidal layer porosity and cake-enhanced osmotic pressure were assumed to be changing in time and were determined simultaneously using a recursive algorithm. 11 Wang and Tarabara [68] proposed a different approach to study the interplay between colloidal fouling and salt concentration polarization. The concentration of rejected salt at the membrane surface when colloidal particles were deposited on the membrane was deterrnined experimentally based on measured salt permeability constant. Then the structure of the fouling layer was estimated in terms of effective porosity. This approach allow ed for a clear identification of individual contributions of concentration polarization and colloidal fouling to the permeate flux decline. It was found that, at a higher solution ionic strength, the effective porosity of the deposit layer was smaller and the effect of colloidal fouling on salt concentration polarization was more evident. 1.2.2- Combined colloidal and organic fouling Organic matter is an important membrane foulant. The effects of hydrodynamics on orgarri c fouling are similar to that on colloidal fouling. In addition, organic fouling is often affected by the foulants’ physicochemical characteristics such as charge, chemical filflCtiOnality, and molecular conformation ([42, 45, 83, 84], and the references therein). Divalent cations, such as Ca2+ and Mg2+ form complexes with organics and greatly increase organic fouling [85]. These factors lead to a selective removal of NOM by membrane, as shown recently by a series of membrane filtration experiments [86]. SCMfer et a1. [41] studied MF membrane filtering surface water containing NOM and hematite colloids. It was found that the solution chemistry (especially pH and calcium 12 {0115) played a very important role in the combined fouling. More specifically, at extreme high or low pH the NOM was absorbed onto the surface of the hematite colloids, which increased the stability of the colloids and thus reduced the membrane fouling. If the colloids pre-aggregated in electrolyte solution prior to adsorption of NOM, the presence of NOM increased the flux decline. Jermann et al. [43] showed that NOM fouling and colloidal fouling (alginate) interacted in a UF membrane filtration system. The fouling from alginate alone was mostly reversible. However, in the presence of NOM, alginate resulted in more severe irreversible fouling. This work implied that substances with a mirror individual influence might have a large impact in the presence of other foulants. Chen et al. [42] used ultrasound to control the combined fouling from NOM and silica Colloi ds in ME membrane filtration. It was found that ultrasound was more effective in enhancing permeate flux under conditions of high pH, low ionic strength, and in the absence of divalent cations. Van der Bruggen at al. [46] demonstrated that the rejection of organic matter (benzyl 31901101 ) by the NF membrane was enhanced in the presence of the polystyrene particles, WhiCh Was more pronounced for smaller particles than for larger particles. This effect was attributed to the adsorption of the organic matter onto the particles. Lee et al. [44] studied the combined fouling in NF membrane filtration with Special attention given to the effect of NC)1\v‘I—calcium complexation. It was found that the flux decline in presence of the combined fouling is less than the sum of the decline due to the individual colloidal and NOM fouling. This effect was presumably attributed to the increased stability of the 13 colloids in the presence of NOM and the competition between the colloids and NOM for calcium. Li and Elimelech [45] demonstrated a synergistic effect between colloidal and organic fouling in a NF filtration system. It was found that the combined fouling resulted in a higher flux decline rate than the sum of the flux decline due to colloidal and organic fouling alone. Solution chemistry and colloidal particle size was found to influence the synergistic effect. 1.2-3- Colloidal feuling and scaling Seal in g from sparingly soluble salts, such as carbonate and sulfate, greatly impedes the application of R0 membranes. Scaling occurs when the solubility of the salts is exceeded. Carbonate scaling can be controlled by decreasing the pH of the feed solution. However Calcium sulfate dihydrate (gypsum) is insensitive to pH, and can not be controlled by adjus ti n g pH. The efi‘ect of crossflow rate and system configuration on gypsum scaling in membrane filtration systems has been extensively examined. Lee and Lee [87] demonstrated that in NF and RO membrane systems the gypsum scaling rate was decreased at higher CfOSSflOw rate and lower transmembrane pressure. Lee and Lueptow [88] used a rotating R0 meIrrbrane system to reduce the gypsum scaling. Generally speaking, these studies achieVed reduction in gypsum crystallization by decreasing the effect of concentration polarization which resulted in lower CaSO4 concentration at membrane surface. 14 It has been found that at high temperature the induction time of gypsum crystallization was decreased [89, 90]. The activation energy of gypsum crystallization can be estimated based on the dependence of induction time on temperature. Cohen’s research group [91, 92] demonstrated progressive axial development of gypsum crystals along the membrane Surf ace, consistent with the increase in concentration polarization. Also, the membrane Surf ace coverage by gypsum increased at higher gypsum saturation index [91, 92]. Sheiklroleslami and co-workers ([93] and references therein) demonstrated that the presence of other salts (such as CaCO3) affected the structure and strength of gypsum crystals. Amjad and Hooley [94] showed that polyelectrolytes affected gypsum Crystallization and the effect depended on the concentration, molecular weight, and Composition of the polyelectrolytes. It is surprising that there is very limited research published on the combined effect of scalin g and colloidal fouling on the membrane performance. Most of the available knOWledge comes from studies on the performance of heat exchangers, where the effect 0f Particles on scaling has been demonstrated to be complex [95]. Bansal et al. [96] found that the effect of suspended particles on scaling depended on the crystallizing ability of pafliCIes. Crystallizing CaSO4 particles greatly enhanced gypsum formation because they provided extra nucleation sites. On the contrary, noncrystallizing alumina particles reduced the gypsum formation. They believed that this effect was due to 1) particles reduced the strength of crystals deposited and thus resulted in a higher crystal removal rate. and 2) particles on crystal surface acted as a distorting agent. Andritsos and 15 Karabelas [97] showed that CaC03 precipitation rate was greatly enhanced by fine aragonite (CaCO3 polymorph) particles, but was not affected by silica colloids. McGarVey and Turner [98] examined the interplay between CaCO3 precipitation and colloidal deposition. It was found although the rate of CaCO3 precipitation was not affected by the silt and hematite (Fe203) particles, the deposition of the particles was emanced by CaCO3 precipitation. Wang and Tarabara [47] studied gypsum scaling and colloidal fouling in a crossflow fr\trati on system. It was found that in the presence of silica colloids, gypsum scaling resulted in a much faster permeate flux decline of the RO membrane, which was not the Simple sum of the flux decline from colloidal deposition and that from gypsum scaling. SEM/ EDS studies demonstrated that colloids attached onto the surface of precipitated gypsum crystals, increasing resistance of the membrane deposit and making hydraulic cleanin g inefficient. 1.2-4- Other combined fouling Studying gypsum crystallization in a nanofiltration (NF) membrane system filtering agfiCUItmal drainage water, Le Gouellec and Elimelech [48] demonstrated that the presence of NOM resulted gypsum crystals of different morphology, which possibly effected the performance of the membrane. It was also found that a single type of commercially available aquatic humic acid could not model the effect of the complex NOM in natural water on gypsum crystallization. 16 1.3. The future research need As is evident from the information presented in Tables 1.1 and 1.2, although combined fouling always occurs, very few studies have been focused on the effect of combined fouling on membrane performance. Systematic studies on combined fouling are needed in order to better predict and control membrane fouling, as which occurs in practice. Future studies can be expected to: (l) Reveal how the individual fouling mechanisms interact, and study the effect of the interactions on membrane performance; (2) Provide guidelines on the optimal strategies for membrane pretreatment and cleaning in specific combined fouling scenarios; and, (3) Develop a membrane fouling index that more accurately predicts fouling propensity of waters of complex compositions. l7 Reference 10. 11. 12. 13. D. Bursill, Drinking water treatment - understanding the processes and meeting the challenges, Wat. Sci. Technol.: Water Supply 1 (2001) 1-7. State of California, Findings and Recommendations, (2003). M. A. Shannon, P. W. Bohn, M. Elimelech, J. G. Georgiadis, B. J. Mariflas and A. M Mayes, Science and technology for water purification in the coming decades, Nature 452 (2008) 301-310. United States Environmental Protection Agency, Report: Membrane filtration guidance manual (815-R—06-009), (2005). AWWA Membrane Technology Research Committee, Committee report: Membrane processes, J. Am. Water Works Assoc. 90 (1998) 91-105. B. S. Kamik, S. H. R. Davies, K. C. Chen, D. R. Jaglowski, M. J. Baumann and S. J. Masten, Effects of ozonation on the permeate flux of nanocrystalline ceramic membranes, Water Res. 39 (2005) 728. J. S. Taylor and MR. Wiesner, Membranes in Water Quality And Treatment. Fifth ed, ed. R.D. Letterman. 1999, New York: McGraw-Hill. 11.1-11.37. S. K. Kang and K. H. Choo, Use of MP and UP membranes for reclamation of glass industry wastewater containing colloidal clay and glass particles, J. Membr. Sci. 223 (2003) 89. M. D. Afonso and R. Borquez, Review of the treatment of seafood processing wastewaters and recovery of proteins therein by membrane separation processes -- prospects of the ultrafiltration of wastewaters from the fish meal industry, Desalination 142 (2002) 29. V. Bonnelye, L. Guey and J. Del Castillo, UF/MF as RO pre-treatment: the real benefit, Desalination 222 (2008) 59. G. Pearce, Introduction to membranes: Manufacturers' comparison: part 1, Filtr. Sep. 44 (2007) 36. M. Barger and RP. Carnahan, Fouling prediction in reverse osmosis processes, Desalination 83 (1991) 3-33. G. Belfort, R. H. Davis and A. L. Zydney, The behavior of suspensions and macromolecular solutions in crossflow microfiltration, J. Membr. Sci. 96 (1995) 1-58. 18 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. MR. Wiesner and S. Chellam, The promise of membrane technology, Environ. Sci. Technol. 33 (17), 360A (1999). S. Lattemann and T. Hopner, Environmental impact and impact assessment of seawater desalination, Desalination 220 (2008) 1. R. W. Baker, Membrane Technology and Applications. 2004, England: McGraw- Hill. Sandia National Laboratories, Desalination and Water Purification Technology Roadmap, Desalination & Water Purification Research & Development Program Report # 95 (2003). C. Gabelich, T. I. Yun, J. F. Green, I. H. Suffet and W. R. Chen, Evaluation of Precipitative Fouling for Colorado River Water: Desalination Using Reverse Osmosis, Desalination and Water Purification Research and Development Program Report No. 85 (2002). K. T. Chua, M. N. A. Hawlader and A. Malek, Pretreatment of seawater: Results of pilot trials in Singapore, Desalination 159 (2003) 225-243. Harvey Winters, Twenty years experience in seawater reverse osmosis and how chemicals in pretreatment affect fouling of membranes, Desalination 110 (1997) 93. Y. R. Ning, L. T. Troyer and S. R. Tominello, Chemical control of colloidal fouling of reverse osmosis systems, Desalination 172 (2005) 1-6. B. G. Darton and M. F azel. A statistical review of 150 membrane autopsies. in International Water Conference. 2001. J. W. McCutchan and J. S. Johnson, Reverse osmosis at Coalinga, California, J. Am. Water Works Assoc. 62 (1970) 346. M. Luo and Z. Wang, Complex fouling and cleaning in-place of a reverse osmosis desalination system, Desalination 141 (2001) 15-22. J. van Agtmaal, H. Huiting, P. A. de Boks and L. L. M. J. Paping, Four years of practical experience with an Integrated Membrane System (IMS) treating estuary water, Desalination 205 (2007) 26. S. F. E. Boerlage, M. Kennedy, M. P. Aniye and J. C. Schippers, Applications of the MFI-UF to measure and predict particulate fouling in R0 systems, J. Membr. Sci. (2003) 97-116. 19 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. A. D. Khawaji, I. K. Kutubkhanah and J. Wie, A 13.3 MGD seawater R0 desalination plant for Yanbu Industrial City, Desalination 203 (2007) 176. Christopher J. Gabelich, Tae I. Yun, Bradley M. Coffey and I. H. Mel Suffet, Pilot-scale testing of reverse osmosis using conventional treatment and microfiltration, Desalination 154 (2003) 207. A. Plottu, B. Houssais, C. Democrate, D. Gate] and J. Cavard, Autopsies of membranes fouled on Mery-sur-Oise pilot units: Many lessons for the behaviour of the water treatment plant, Desalination 157 (2003) 367. L. Y. Dudley and E. G. Darton, Membrane autopsy -- a case study, Desalination 105 (1996) 135 -141. C. J. Gabelich, M. D. Williams, A. Rahardianto, J. C. Franklin and Y. Cohen, High-recovery reverse osmosis desalination using intermediate chemical demineralization, J. Membr. Sci. 301 (2007) 131. A. Boubakri and S. Bouguecha, Diagnostic and membrane autopsy of Djerba Island desalination station, Desalination 220 (2008) 403-411. M. Karime, S. Bouguecha and B. Hamrouni, RO membrane autopsy of Zarzis brackish water desalination plant, Desalination 220 (2008) 258 - 266. F. H. Butt, F. Rahman and U. Baduruthamal, Characterization of foulants by autopsy of RO desalination membranes, Desalination 114 (1997) 51 - 64. H. L. Yang, C. Huang and J. R. Pan, Characteristics of RO foulants in a brackish water desalination plant, Desalination 220 (2008) 353-358. A. S. Al-Amoudi and A. M. Farooque, Performance restoration and autopsy of NF membranes used in seawater pretreatment, Desalination 178 (2005) 261 - 271. S. Belfer, J. Gilron, Y. Purinson, R. Fainshtain, N. Daltrophe, M. Priel, B. Tenzer and A. Toma, Effect of surface modification in preventing fouling of commercial SWRO membranes at the Eilat seawater desalination pilot plant, Desalination 139 (2001) 169 - 176. L. D. Nghiem and A. I. Schafer, Fouling autopsy of hollow-fibre MF membranes in wastewater reclamation, Desalination 188 (2006) 113 - 121. J. Kim, F. A. DiGiano and R. D. Reardon, Autopsy of high-pressure membranes to compare effectiveness of MP and UP pretreatment in water reclamation, Wat. Research 42 (2008) 697 - 706. 20 40. 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. R. Y. Ning and T. L. Troyer, Colloidal fouling of RO membranes following MF/U F in the reclamation of municipal wastewater, Desalination 208 (2007) 232. A. I. Schafer, U. Schwicker, M. M. Fischer, A. G. Fame and T. D. Waite, Microfiltration of colloids and natural organic matter, J. Membr. Sci. 171 (2000) 151-172. D. Chen, L. K. Weavers, H. W. Walker and J. J. Lenhart, Ultrasonic control of ceramic membrane fouling caused by natural organic matter and silica particles, J. Membr. Sci. 276 (2006) 135-144. D. Jermann, W. Pronk, S. Meylan and M. Boller, Interplay of different NOM fouling mechanisms during ultrafiltration for drinking water production, Water Res. 41 (2007) 1713-1722. S. Lee, J. Cho and M. Elimelech, Combined influence of natural organic matter (NOM) and colloidal particles on nanofiltration membrane fouling, J. Membr. Sci. 262 (2005) 27-41. Q. Li and M. Elimelech, Synergistic effects in combined fouling of a loose nanofiltration membrane by colloidal materials and natural organic matter, J. Membr. Sci. 278 (2006) 72. B. Van der Bruggen, J. H. Kim, F. A. DiGiano, J. Geens and C. Vandecasteele, Influence of MP pretreatment on NF performance for aqueous solutions containing particles and an organic foulant, Sep. Purif. Technol. 36 (2004) 203. P. Wang and V.V. Tarabara. Coupled effects of scaling and colloidal deposition on the performance of reverse osmosis membranes. in The 233rd ACS National Meeting. March 2007. Chicago, IL. Y. A. Le Gouellec and M. Elimelech, Calcium sulfate (gypsum) scaling in nanofiltration of agricultural drainage water, J. Membr. Sci. 205 (2002) 279. S. Chellam and M. R. Wiesner, Particle transport in clean membrane filters in laminar flow, Environ. Sci. Technol. 26 (1992) 1611-1621. W. Richard Bowen, Anne Mongruel and Paul Williams, Prediction of the rate of cross-flow membrane ultrafiltration: A colloidal interaction approach, Chem. Eng. Sci. 51 (1996) 4321-4333. S. Hong, R. S. Faibish and M. Elimelech, Kinetics of permeate flux decline in crossflow membrane filtration of colloidal suspensions, J. Colloid Interface Sci. 196 (1997) 267. 21 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. G. Belfort, R.H. Davis and AL. Zydney, The behavior of suspensions and macromolecular solutions in crossflow microfiltration, J. Membr. Sci. 96 (1994) 1-58. P. Bacchin, P. Aimar and R. W. Field, Critical and sustainable fluxes: Theory, experiments and applications, J. Membr. Sci. 281 (2006) 42. E. M. Hock, S. Hong and M. Elimelech, Influence of membrane surface properties on initial rate of colloidal fouling of reverse osmosis and nanofiltration membranes, J. Membr. Sci. 188 (2001) 115. X. Zhu and M. Elimelech, Colloidal fouling of reverse osmosis membranes: measurements and fouling mechanisms, Environ. Sci. Technol. 31 (1997) 3654- 3662. K. Boussu, A. Belpaire, A. Volodin, C. Van Haesendonck, P. Van der Meeren, C. Vandecasteele and B. Van der Bruggen, Influence of membrane and colloid characteristics on fouling of nanofiltration membranes, J. Membr. Sci. 289 (2007) 220. B. Van der Bruggen, D. Segers, C. Vandecasteele, L. Bracken, A. Volodin and C. Van Haesendonck, How a microfiltration pretreatment affects the performance in nanofiltration, Sep. Sci. Technol. 39 (2005) 1443-1459. K. Boussu, C. Vandecasteele and B. Van der Bruggen, Relation between membrane characteristics and performance in nanofiltration, J. Membr. Sci. 310 (2008)51. A. V. R. Reddy and H. R. Patel, Chemically treated polyethersulfone/polyacrylonitrile blend ultrafiltration membranes for better fouling resistance, Desalination 221 (2008) 318. R. M. McDonogh, C. J. D. Fell and A. G F ane, Surface charge and permeability in the ultrafiltration of non-flocculating colloids, J. Membr. Sci. 21 (1984) 285- 294. K.-J. Hwang, H.-C. Liu and W.-M. Lu, Local properties of cake in cross-flow microfiltration of submicron particles, J. Membr. Sci. 138 (1998) 181-192. Y. Lee and M. M. Clark, Modeling of flux decline during crossflow ultrafiltration of colloidal suspensions, J. Membr. Sci. 149 (1998) 181-202. P. Harmant and P. Aimar, Coagulation of colloids in a boundary layer during cross-flow filtration, Colloids Surf, A 138 (1998) 217-230. 22 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. V. V. Tarabara, I. Koyuncu and M. R. Wiesner, Effect of hydrodynamics and solution chemistry on permeate flux in crossflow filtration: Direct experimental observation of filter cake cross sections, J. Membr. Sci. 241 (2004) 65-78. J. C. Chen, M. Elimelech and A. S. Kim, Monte Carlo simulation of colloidal membrane filtration: Model development with application to characterization of colloid phase transition, J. Membr. Sci. 255 (2005) 291-305. M. R. Wiesner, Morphology of particle deposits, J. Environ. Eng. ASCE 125 (1999)1124-1132. R. S. Faibish, M. Elimelech and Y. Cohen, Effect of interparticle electrostatic double layer interactions on permeate flux decline in crossflow membrane filtration of colloidal suspensions: An experimental investigation, J. Colloid Interface Sci. 204 (1998) 77-86. F. Wang and V.V. Tarabara, Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance, J. Membr. Sci. 293 (2007) 111-123. A. E. Childress and M. Elimelech, Effect of solution chemistry on the surface charge of polymeric reverse osmosis and nanofiltration membranes, J. Membr. Sci. 119 (1996) 253. J. J. Qin, M. H. 00 and F. S. Wong, Effects of pH and antiscalant on fouling of RO membrane for reclamation of spent rinse water from metal plating, Sep. Purif. Technol. 46 (2005) 46. K. J. Hwang and C. L. Hsueh, Cross-flow microfiltration of aggregated submicron particles, J. Membr. Sci. 201 (2002) 137-148. H. Vyas, A. J. Mawson, R. .1. Bennett and A. D. Marshall, A new method for estimating cake height and porosity during crossflow filtration of particulate suspensions, J. Membr. Sci. 176 (2000) 113-119. L. J. Thibodeaux, Chemodynamics: Environmental Movement of Chemicals in Air, Water, and Soil. 1979: John Wiley & Sons. B. P. Boudreau, The diffusive tortuosity and porosity of fine-grained sediments, Geochim. Cosmochim. Acta 60 (1996) 3139-3142. P. C. Carman, Fluid flow through a granular bed, Trans. Inst. Chem. Eng. London 15(1937)150-166. P. Lipp, R. Gimbel and F. H. Frimmel, Parameters influencing the rejection properties of FT30 membranes, J. Membr. Sci. 95 (1994) 185-197. 23 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. J. M. Jackson and D. Landolt, About the mechanism of formation of iron hydroxide layers on reverse osmosis membranes, Desalination 15 (1973) 361-378. E. M. Hoek, A. S. Kim and M. Elimelech, Influence of crossflow membrane filter geometry on colloidal fouling of reverse osmosis membranes, Environ. Eng. Sci. 19 (2002) 357-372. E. M. Hoek and M. Elimelech, Cake-enhanced concentration polarization: 3 new fouling mechanism for salt-rejecting membranes, Environ. Sci. Technol. 17 (2003) 5581-5588. S. Lee, J. Cho and M. Elimelech, Influence of colloidal fouling and feed water recovery on salt rejection of R0 and NF membranes, Desalination 160 (2004) 1- 12. A. I. Schafer, A. G. F ane and T. D. Waite, Fouling effects on rejection in the membrane filtration of natural waters, Desalination 131 (2000) 215-224. H. Y. Ng and M. Elimelech, Influence of colloidal fouling on rejection of trace organics contaminants by reverse osmosis, J. Membr. Sci. 244 (2004) 215-226. M. Taniguchi, J. E. Kilduff and G. Belfort, Modes of natural organic matter fouling during ultrafiltration, Environ. Sci. Technol. 37 (2003) 1676 - 1683. A. W. Zularisam, A. F. Ismail and R. Salim, Behaviours of natural organic matter in membrane filtration for surface water treatment - a review, Desalination 194- 231 (2006) 211. A. I. Schafer, A. G. F ane and T. D. Waite, Nanofiltration of natural organic matter: Removal, fouling and the influence of multivalent ions, Desalination 118 (1998) 109. J. Kim, W. Shi, Y. Yuan and M. M. Benjamin, A serial filtration investigation of membrane fouling by natural organic matter, J. Membr. Sci. 294 (2007) 115. S. Lee and CH. Lee, Effect of operating conditions on CaSO4 scale formation mechanism in nanofiltration for water sofiening, Water Res. 34 (2000) 3854. S. Lee and RM. Lueptow, Control of scale formation in reverse osmosis by membrane rotation, Desalination 155 (2003) 131. Fathi Ahmi and Abdellatif Gadri, Kinetics and morphology of formed gypsum, Desalination 166 (2004) 427. 24 90. 91. 92. 93. 94. 95. 96. 97. 98. A. Lancia, D. Musmarra and M. Prisciandaro, Measuring induction period for calcium sulfate dihydrate precipitation, AIChE J. 45 (1999) 390 - 397. W.Y. Shih, A. Rahardianto, R.W. Lee and Y. Cohen, Morphometric characterization of calcium sulfate dihydrate (gypsum) scale on reverse osmosis membranes, J. Membr. Sci. 252 (2005) 253-263. A. Rahardianto, W.Y. Shih, R.W. Lee and Y. Cohen, Diagnostic characterization of gypsum scale formation and control in RO membrane desalination of brackish water, J. Membr. Sci. 279 (2006) 655. R. Sheikholeslami, Mixed salts -- scaling limits and propensity, Desalination 154 (2003) 117. Z. Amjad and J. Hooley, Influence of polyelectrolytes on the crystal growth of calcium sulfate dihydrate, J. Colloid Interface Sci. 111 (1986) 496. D. Hasson. Progress in precipitation fouling: a review, in:Understanding Heat Exchanger Fouling and its Mitigation, T.R. Bott et al. (Eds). 1999: Begell House, New York. B. Bansal, H. Muller-Steinhagen and X. D. Chen, Effect of suspended particles on crystallization fouling in plate heat exchangers, J. of Heat Transfer, Transactions ASME 119 (1997) 568. N. Andritsos and A.J. Karabelas, The Influence of particulates on CaCO3 scale formation, J. Heat Transfer 121 (1999) 225-227. G.B. McGarvey and CW. Turner, Synergism models for heat exchanger fouling mitigation. In: T.R. Bott et al.Understanding Heat Exchanger Fouling and its Mitigation, Begell House, New York, (1999) 155-161. 25 CHAPTER TWO Wang, Fulin; Tarabara, Volodymyr V. (2007). Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance. J. Membr. Sci. 293 111—123 26 CHAPTER TWO COUPLED EFFECTS OF COLLOIDAL DEPOSITION AND SALT CONCENTRATION POLARIZATION ON REVERSE OSMOSIS MEMBRANE PERFORMANCE ABSTRACT Concentration of rejected salt at the membrane surface when colloidal particles were deposited on the membrane was determined experimentally based on measured salt permeability constant. This approach allowed for a clear identification of individual contributions of concentration polarization and colloidal fouling to the permeate flux decline. The observed transient behavior of permeate flux, salt rejection, mass of deposited colloids, and effective porosity of the deposited layer unequivocally pointed to the importance of the two-way coupling between salt concentration polarization and colloidal deposition. After the initial increase in the effective porosity of the fractal colloidal deposit, the porosity reached a maximum, and gradually decreased to a steady state value that depended strongly on the solution ionic strength. Both the resistance of the colloidal deposit and osmotic pressure, which could be enhanced in the presence of colloids, contributed significantly to the permeate flux decline. A short term increase in salt rejection was observed upon introduction of colloidal particles into the feed in the absence of membrane channel spacer. This increase could be sustained over the long term in the presence of spacers and was attributed to the depolarization of the salt boundary layer due to the local mixing by the concentrated flowing layer of colloidal particles. 27 2.1. INTRODUCTION The shrinking water supply together with our continuously increasing demand for this resource has stimulated a renewed interest in reverse osmosis (RO) as a water treatment unit process. Desalination, softening, disinfection by-product control, and removal of specific inorganic contaminants such as arsenic, barium, nitrate, nitrite and others exemplify possible applications of R0 in water treatment. A wider use of RO membranes, however, is impeded by a group of phenomena that lead to permeate flux decline, deterioration of permeate water quality, shortening of membrane life, and thus to a substantial increase in operational costs [1, 2]. The relevant physicochemical phenomena include concentration polarization, scaling, adsorptive and biological fouling, and colloidal fouling. Although source waters usually undergo pretreatment that is designed to remove the colloidal fraction, in practice both dissolved and residual suspended phases are present in R0 feed [3-5]. The origins, surface chemistry, concentration, and morphology of the smaller colloids that persist through the pretreatment stage are likely to depend on the water source and the type of pretreatment. Examples of colloidal matter in R0 feed waters include elemental sulfur, silica, and ferric and aluminum hydroxides [3-7]. Because of their small size, these residual colloids can form membrane deposits with a high hydraulic resistance contributing to the permeate flux decline. Also, colloids deposited on the membrane surface can inhibit back-diffusion of dissolved species resulting in precipitation of sparingly soluble salts in the pores of the deposited layer and 28 in the enhanced osmotic pressure [8, 9]. If the source water pretreatment is inadequate or fails, colloidal fouling can be especially egregious. Given that practical applications of R0 usually involve multicomponent feeds and because RO membranes provide an absolute barrier to most components of the feed water, potential interactions of colloids and other rejected materials have to be taken into account in predicting membrane performance and deciding on the optimal strategy for membrane cleaning [3, 8]. Fouling of RO membranes by iron hydroxide [IO-I3], polystyrene [l4], and silica [9, 15- 17] colloids has also been investigated in controlled laboratory studies. In all cases, permeate flux was reported to decrease with the introduction of colloids into the feed. Another important metric of membrane performance, solute rejection, has been shown to be influenced by the presence of colloidal layer on the membrane surface. Rejection of NaCl by RC membranes was shown to decrease in the presence of membrane fouling by iron hydroxide [10, 18] and by silica colloids [9, 16, 18, 19]. Recently, Hoek et al. have proposed a model that described how hindrance of the back-diffusion of salt by a stagnant colloidal cake results in cake-enhanced concentration polarization, an increase in osmotic pressure, and, consequently, lower fluxes and salt rejections [9, 16]. Results of the model, which assumed the porosity of the colloidal silica layer to be constant, agreed well with experimental data on permeate flux and salt rejection. A modified cake-enhanced concentration polarization model was later proposed by Ng and Elimelech [17], wherein the colloidal layer porosity and cake-enhanced osmotic pressure were assumed to be changing in time and were determined simultaneously using a recursive algorithm. 29 In experiments with other colloids or colloidal mixtures, the rejection of salt has been reported to exhibit more complex behavior. RO membrane fouling by humic substances alone has been shown to result in a slight increase [20] or a marginal decrease in NaCl rejection [21]. However, when silica colloids were present in the feed, the rejection of NaCl was shown to remain unchanged with respect to that when no fouling occurred [21]. Similarly, for suspensions containing iron oxide colloids and humic substances, the rejection of NaCl was shown to improve with respect to NaCl rejection observed when humic substances were the sole foulant [20]. In a study of effects of nanofiltration membrane fouling on rejection, Shafer et al. [22] reported that rejection of cations and negatively charged low molecular weight acids depended on the charge of ferric chloride precipitate deposited on the membrane surface. In case of micro- and ultrafiltration, the accumulation of larger colloids on the membrane surface is also known to modify rejection of smaller colloids and solutes. For example, rejection of dextran was shown to decrease in the presence of bentonite deposit on the surface of rrricrofiltration membrane and was attributed to retention of dextran molecules in the cake layer and in membrane pores [23]. In another study on the microfiltration of colloidal mixtures, the retention of 50 nm gold colloids was significantly reduced in the presence of a membrane deposit of 1 pm latex colloids [24]. Improvement of solute rejection in the presence of dynamically formed layers (secondary or dynamic membranes) of different types of particles on membrane surfaces has also been reported ([25] and references therein). These observations point to a strong dependence of membrane rejection on the type of the foulants involved and, more generally, on the properties of the fouling layer. 30 The structure of the colloidal fouling layer is known to be a complex function of operational variables such as flow hydrodynamics, colloid size and charge, and solution chemistry [26-31]. The propensity of colloids to deposit affects the layer’s hydraulic resistance both directly via the overall deposit thickness and indirectly via the accumulative drag force on deposited colloids that plays a role in determining the equilibrium distance between colloidal particles. Therefore, accounting for various mechanisms of particle transport to membrane surface in crossflow channelsiis essential for the accurate prediction of the permeate flux [32]. It should be noted that the hydrodynamics of colloids in membrane channels can be drastically modified in the presence of spacers. It is surprising that despite the wide application of spacers in practice [33], no information is available on their effect on the colloidal fouling of RO membranes. Solution chemistry-mediated interactions between particles within the deposited layer is another significant factor that determines the structure and, therefore, specific hydraulic resistance of the deposit [34]. Theoretical and experimental studies indicate that colloidal cakes can be stratified and can contain layers with dense and loose particle packing [27, 29, 30, 35]. There is also evidence for the existence of the concentrated flowing layer of colloids between the stagnant cake of deposited colloids and the bulk suspension [36-39]. One objective of this study is the direct experimental determination of (I) the effect of rejected salt on the structure of colloidal layer and (2) how the two-way coupling between salt concentration polarization and colloidal deposition influences permeate flux and salt 31 rejection. Another objective is the evaluation of the effect of spacers on RO flux and rejection in the presence of colloidal particles in the feed. 2.2. BACKGROUND Reverse osmosis membrane performance is characterized in terms of permeate flux and solute rejection. When the feed water is free of solutes and potential foulants, the pure water permeate flux J 0 is linked to the transmembrane pressure differential AP by the Darcy’s equation: J0 = 9 (2-1) where Rmo is the hydraulic resistance of the membrane to pure water, and ,u is the dynamic viscosity of water. When a solute is present in the feed and there is a gradient of solute concentration across the membrane, osmotic pressure (Arrm) must be overcome. In the simple case of dilute 1-1 electrolyte, van’t Hoff equation can be used to calculate Air", : Mm = 2RT(Cm—Cp), (2.2) where R is the universal gas constant, T is the absolute temperature, Cm is the solute concentration at the membrane surface, and C p is the solute concentration in the 32 permeate. Due to the effect of concentration polarization, solute concentration at the surface of solute-rejecting membranes is higher than solute concentration in the bulk of the feed solution. The boundary layer film model can be used to estimate C m [40]: 92.—fez“ (£]_ex (1] t 23 Cf-Cp p D P k , (-) where C f is the solute concentration in the bulk of the feed solution, J is the volumetric permeate flux, D is the solute diffusion coefficient in water, 6 is the concentration . . . D . . polarization boundary layer thickness, and k = 5 IS the mass transfer coeffrcrent. For laminar flow in a flat rectangular channel, k can be estimated from Sherwood correlation [41]: 1/3 Sh=-k%h—=(3.663 +1.613 °Re-Sc--quZ-] , (2.4) where Sh is the Sherwood number, L is the channel length , d}, is the hydraulic diameter (d), z2h) of the membrane channel and h is the channel height, Re is the Reynolds number, and Sc is Schmidt number. Equation (2.4) holds when the condition 0.1<(Re-Sc-i:—)<104 is satisfied. Equations (2.2) and (2.3) can be combined to produce an expression for Arm: 33 J Alrm = ZRTCfRObs emf/2), (25) where Robs is the observed solute rejection: C Rob, =1- C?" (2.6) When colloidal foulants are present in the feed and are rejected by the membrane, they may accumulate on the membrane surface and form a deposit that exerts hydraulic resistance (Rd) to the permeate flux and contributes to its decline. Also, the membrane resistance (Rm) tends to increase measurably during the conditioning stage as the membrane is exposed to the electrolyte solution free of foulants. As will be shown in section 4.2, it becomes important to account for this difference between Rm and Rmo when estimating osmotic pressures at lower ionic strengths. When these three factors - Azrm , Rd , and (Rm —R,,,0) - are accounted for and the membrane reflection coefficient is assumed to be equal to unity, the expression for water flux across the membrane becomes: __ AP— An," ”(Rm + Rd) (2.7) 34 The interaction between the deposited layer and rejected salts makes it difficult to factor out individual contributions of Arr,n and Rd to the flux decline. On one hand, in the presence of the fouling layer, correlations such as Eq. (2.4) for the mass transfer coefficient k are not applicable and, therefore, Eq. (2.5) can not be used to compute Cm and Arm. On the other hand, Rd depends on the deposit’s mass and porosity that are functions of the ionic strength. Thus, without experimentally measuring one of the two unknowns (Azz'm or Rd ), it is not possible to experimentally determine separate effects of osmotic pressure and colloidal deposition on the membrane performance. This dilemma can be resolved by measuring the membrane salt permeability constant, B . According to the solution-diffusion model [42], solute flux J s can be related to Cm and Cas p J, = B(C,,, — C p). (2.8) The mass solute flux J 5 (Eq. (2.8)) and the volumetric water flux J (Eq. (2.7)) are related via the concentration of salt transported across the membrane: JC Ks ca II p. (29) Equations (2.8) and (2.9) can be combined to determine B when colloidal foulants are not present in the feed. As will be shown in the section 2.4.3, B was found to be 35 approximately constant in the experiments reported in this paper. This allowed us to use the measured value of B to determine Air", when colloids were deposited on the membrane surface. By combining Eqs. (2.2), (2.8), and (2.9) we get: JCp Air," = 2RT(Cm—Cp) = 2RT -—B— , (2.10) where C p and J are determined experimentally. After Art", is determined from Eq. (2.10), Eq. (2.7) can be used to compute Rd. Kozeny-Carman equation [43] for Rd can be used then to evaluate the effective porosity e of the deposited layer: 1801-5 M Rd= (23) P, (2.11) ppdpe where A is the membrane filtration area, M p is the mass of particles deposited on membrane surface, and dp and ,0p are the diameter and the density of the silica particles, respectively. 36 2.3. EXPERIMENTAL METHODS 2.3.1. Experimental apparatus The schematic of the bench-scale crossflow RO unit is shown in Fig. 2.1. Two identical Sepa CF 11 high pressure membrane modules in medium/high foulant configuration (GE Osmonics, Minnetonka, MN) were connected in parallel. Positive displacement pump (M03 Hydracell, Wanner Engineering, Minneapolis, MN) equipped with a flowrate control unit (KBMD-240D, KB Electronics Inc., Coral Springs, FL) was used to deliver feed water to the filtration cell. Pulsation dampener (Model H1020V, Blacoh Fluid Control Inc., Riverside, CA) was installed immediately downstream from the pump outlet. Back pressure regulator (BP-3, Circle Seal Controls Inc., Corona, CA) was used to maintain the transmembrane pressure differential at a constant predetermined value in the range of 150 psi (1.034 MPa) to 500 psi (3.477 MPa); the pressure was continuously monitored by an analog pressure transducer (PX303-500G5V, Omega, Stamford, CT). In-line digital flow meters (Models 101-8 and 101-3, McMillan Co., Georgetown, TX) were used to continuously monitor the retentate rate and permeate flux. The meters were calibrated using a digital weighing balance (ARC120, Ohaus Corp., Pine Brook, NJ) interfaced with a computer via a built-in RS-232 port. The permeate was collected to a vessel positioned on the balance, as indicated by thick dashed lines in Fig. 2.1, only when determining water and salt permeability constants. In all other experiments, the permeate was recirculated back into the feed tank. In all experiments, the retentate flow rate was kept at 4.8 L/min, which resulted in a crossflow velocity of 0.24 m/s and a Reynolds number of 816. 37 The membrane channel mesh spacers (Model YCAFSOS705, GE Osmonics, Minnetonka, MN) were used in several experiments to study the effect of spacers on permeate flux and rejection. Cross section of the mesh spacer elements was approximately rectangular with the height of ca. 1.12 mm and the width of ca. 0.56 mm. Flow-through conductivity probes interfaced with a conductivity meter (Orion 550, Thermo Electron Corp., Beverly, MA) were installed in the permeate line. One conductivity probe (Orion, 013060A, Thermo Electron Corp.) with a flow-through cell was used to measure conductivity in the range of 0.001 uS/cm to 300 uS/cm; another conductivity probe (Model 013005A, Thermo Electron Corp.) was equipped with a house-made glass flow-through cell and was used to measure conductivity in the range of 1 uS/cm to 200 mS/cm. The conductivity cells had a hydraulic detention time that ranged from 0.5 min to 1 min, depending on the value of the permeate flux. During membrane fouling experiments, the permeate conductivity was recorded every 2 min during the first 4 h of the experiment and every 6 minutes afterwards. The data from the pressure transducers, the weighing balance (when used), the conductivity meter, and the flow meters were logged to the computer via the data acquisition module (PCI-6023E/SC-2345, National Instruments Corp., Austin, TX) using 38 _/\/_ system LabVlew data acquisition digital balance 'C--.‘-----.-.---.---.-DOD-COCOODODOO-OC--------.------.0------------ J l .. .m .H. IIIIIIIIIIIIIII J rm mm ....... H mm mm mm nnnnn . . . . .. .. o . u .. "m mm w" n" mumm" fl. mm "ms-mu fl ~H~r FEE mm ”.r rrrrrrrrrrr Lstr rrrrrrr . mm m m m n" " um " "m m mm - Mm I m Pu vmm "u N am. w..." "m u .- emu" "u .::i.b®l man "u mr c. m" mm m .m mm m m" flowmeter modules I—I ID: RO membrane pulse dampener lIllI] external tempeéure sensor um r-—--coco-no-Q-ouocoucoo-o-u-o-o-oo-uu‘-------o-u-o-o-u-u-u-u-uu—uuu II cooling liquid line Feed tank 1. Schematic of the experimental apparatus Figure 2 39 a program written in LabView (version 7.1, National Instruments). These data were acquired at a 30 s time interval and were smoothed using the moving average procedure. The temperature of the feed water in the 28.4 L (7.5 US gal) high density polyethylene feed tank (Nalgene Labware, Rochester, NY) was maintained at (20.0 i 0.5) °C using a programmable circulating chiller (Model 9512, PolyScience, Niles, IL) with an external temperature probe. The chiller circulated the cooling liquid through a custom-made stainless steel (SS316) chilling element immersed in the feed tank. The cooling liquid was a 1:1 (volumezvolume) mixture of distilled water and ethylene glycol. The temperature of the feed suspension was recorded and permeate flux was adjusted to account for the change in viscosity with temperature. The concentration of Silica particles in the feed tank was monitored spectrophotometrically (Multi-Spec 1501, Shimadzu, Kyoto, Japan) by diverting a small portion of the retentate into a flow-through UV-vis sample cell (Model 178.710 - QS, Hellma GmbH & Co KG, Milllheim) for absorption measurement before returning this sample into the feed tank. After each experiment, the crossflow apparatus was cleaned with a detergent solution and then flushed with deionized water four times. 2.3.2. Reagents and chemical solutions 40 All reagents were of ACS analytical grade or higher purity (Fisher Scientific, Pittsburgh, PA) and were used without further purification. The ultrapure water used in the experiments was supplied by a commercial ultrapure water system (Lab Five, USFilter Corp., Hazel Park, MI) equipped with a terminal 0.2 pm capsule microfilter (PolyCap, Whatman P1c., Sanford, ME). The resistivity of water was greater than 16 MQ-cm. The pH of water was monitored (pH meter D12, Horiba, Kyoto, Japan) and was in the range of 6.4 to 6.8 when in equilibrium with atmosphere. No buffers were added to the feed water. 2.3.3. Silica colloids and R0 membranes Silica particles (SnowTex-XL, Nissan Chemical America Corp., Houston, TX) were chosen as the model colloidal foulant. Zeta potential measurements for the silica colloids at several ionic strengths were performed using the zeta potential analyzer (ZetaPALS, Brookhaven Instrument Corp., Holtsville, NY). Before zeta potential measurements, the pH and the ionic strength of the solution were adjusted using KOH, HCl, and KCl. Dynarrric light scattering (Bl-MAS particle sizing module, ZetaPALS, Brookhaven Instrument Corp., Holtsville, NY) was used to measure particle size distribution. Samples were diluted in a 10 mM/L KCI solution to reach the recommended count rate range of 50 kcps to 500 kcps. Transmission electron microscopy (JEOL 100 CX II TEM operated at 100 kV) was used to verify light scattering data. 41 Polyamide thin-film composite RO membrane (BW30-365, Dow-FilmTec, Minneapolis, MN) was used in all experiments. The membrane was supplied as a flat sheet and was stored in a sealed plastic container at room temperature. Steaming potential analyzer (BI- EKA, Brookhaven Instrument Corp., Holtsville, NY) was used to measure streaming potential of membranes. Before each experiment, two rectangular membrane coupons were cut out of the membrane sheet and soaked in deionized water for 24 h at 4 °C with water exchanged after the first 12 h of storage, as recommended by the membrane manufacturer. 2.3.4. Experimental protocol Table 2.1 describes the chronological sequence of experimental tasks. In what follows, the tasks are discussed in detail. 2.3.4.1. Membrane compaction It is well known that most RO membranes exhibit limited and largely irreversible compaction behavior [40]. At the beginning of each experiment, deionized water was filtered through the membranes to ensure that the irreversible compaction would not contribute to the flux decline observed in filtration experiments. Optimization of the compaction procedure involved choosing both compaction transmembrane pressure differentials and compaction time. Pressure at the compaction stage was chosen to exceed pressure to be used in the following membrane fouling experiment. Due to some scatter in the hydraulic resistance of clean membrane coupons and different osmotic pressures in different experiments, the transmembrane pressure differential in filtration tests was 42 adjusted to ensure the same initial permeate flux in all experiments. Pressures used in filtration experiments with deionized water and 10‘2 M, and 10‘1 M NaCl aqueous solutions as feeds were in the range of 250 psi (1.724 MP3) to 400 psi (2.758 MPa). The deionized water had ionic strength of 610'6 M as determined from the standard curve for the conductivity of NaCl solutions. Prior to experiments with these feed waters, the membranes were compacted at transmembrane pressure differentials of 300 psi (2.068 MP3), 350 psi (2.413 MPa), and 450 psi (3.103 MPa), correspondingly. A flux change of less than 1% over 12 h of the continuous compaction was chosen as the quantitative criterion of the complete compaction. It was found that 48 h were sufficient to meet the criterion in all experiments. The compaction was performed with the deionized water that was exchanged once during the compaction. 2.3.4.2. Pure water permeate flux and measurement of membrane hydraulic resistance After compaction, permeate flux was recorded for a series of transmembrane pressure differentials applied: 250 psi (1.723 MPa), 300 psi (2.068 MPa), 350 psi (2.413 MPa), 400 psi (2.758 MPa) and 450 psi (3.103 MPa). The hydraulic resistance of the clean membrane (Rmo) was calculated from Eq. (2.1) using linear least squares fitting. 2.3.4.3. Membrane conditioning After membrane hydraulic resistance was determined, sodium chloride was added into the feed tank to condition the membrane with the solution of the same ionic strength as 43 was to be used in the following filtration experiment. After 12 h of conditioning, stable salt rejection and permeate flux were achieved. The quantitative criterion of the complete conditioning was that the change in flux was less than 0.5% over 2 h of the continuous conditioning. At the end of conditioning, the transmembrane pressure differential was adjusted to achieve the desired initial flux. In our experiments, the initial flux was chosen to be in the 54 L/(mz-h) to 56 L/(mz-h) (1.50-10'5 m/s to 1.55-10'5 m/s) range for all . filtration experiments. This ensured similar initial hydrodynamic conditions for fouling experiments with different membranes. It should also be noted that this value of the initial permeate flux is higher than values of permeate flux commonly encountered in practice (up to 30 L/(mz-h)). This relatively high value was reported in other recent studies on colloidal fouling of RO membranes [9, 16, 17, 19, 21], where similar permeate fluxes were employed. The consequence of this choice of initial permeate flux was that during the initial stage of filtration (a) larger mass of colloids was deposited onto the membrane and b) more significant concentration polarization occurred than one expects for fluxes employed in R0 practice. After the steady state was achieved, however, the value of permeate flux in our experiments was comparable to values observed in practice. 2.3.4.4. Membrane fouling experiments At the end of the membrane conditioning stage, a measured amount of silica particle stock suspension was added into the 25 liters of the feed suspension to achieve the concentration of silica particles of 200 mg/L. The pH of the Silica suspension during the fouling experiments was between 7.1 and 7.9 and was not adjusted. The fouling 44 experiments were stopped after 24 h of filtration, which was sufficient for the steady state permeate flux to be established. The membrane coupons were removed from the membrane cell and stored at 4 °C for future microscopy analyses. To determine the mass of silica particles deposited on the membrane, absorbance of the suspension in the retentate line was continuously recorded at a 2 nrin time interval. The concentration of silica particles were calculated based on the absorption of the suspension at 220 nm using a standard curve recorded beforehand ( ,02 = 1). This wavelength was chosen because of the increasing absorption by NaCl at wavelengths lower than ca. 210 nm and decreasing absorption of silica particles with increasing wavelength. A simple mass-balance equation was applied to calculate the mass of particles deposited on membrane surface (M p ). 2.4. RESULTS AND DISCUSSION 2.4.1. Characteristics of colloidal particles and the membrane The density of colloidal silica particles was calculated gravimetrically to be 2.36 g/cm3. Manufacturer-supplied data indicated that the silica particle diameter ranged from 45 nm to 60 nm. Light scattering data showed that the particle size distribution was relatively narrow (polydispersity factor of 0.074) with a mean hydrodynamic particle diameter of (55.6 d: 1.3) nm. This was close to the value of particle diameter of (51.8 i: 2.4) nm determined by TEM imaging. Because TEM data were less representative of the entire 45 particle population, the particle diameter of 55.6 nm was used in all calculations. Zeta potential values were measured to be (-35.1 i 0.7) mV, (-28.7 i: 1.9) mV, (-25.4 :t 1.5) mV, and (-20.9 :1: 2.1) mV for solution ionic strengths of 6106 M, 1-10'3 M, 1-10'2 M, and 1:101 M, respectively. The general trend for zeta potential of silica particles to become less negative with increasing ionic strength was attributed to the compression of the ionic double layer surrounding particles. The contact angle for the BW30-365 membrane with deionized water was measured to be (72.4 :1: 2.7) °C. The average hydraulic resistance of the membrane was (11.4 :1: 0.3)-10l3 ml (95% confidence interval and p2 > 0.998 for all measurements). Streaming potential of the membrane was measured to be (-16.6 i 0.4) mV at pH 7.2 in the 1 mM KCl electrolyte. The minimal rejection reported by the manufacturer was 99% when measured for 2 g/L NaCl electrolyte at 225 psi (15.3 bar) transmembrane pressure differential, 25°C, pH 8 and 15% recovery. 2.4.2. Changes in membrane resistance during conditioning. Correction factor During the compaction stage, membrane resistance gradually increased by ca. 20% before stabilizing after approximately 50 h of compaction (Fig. 2.2). During the conditioning stage (also known as pretreatment stage [44]), the permeate flux dropped initially with the introduction of NaCl electrolyte into the feed. The magnitude of this decrease could be accurately predicted based on the calculated value of the osmotic pressure due to the introduced electrolyte. Following this initial step-wise decrease, 46 another gradual decline (ca. 3%) in permeate flux was observed that lasted approximately 2 h before the flux stabilized; this decline was observed consistently in all conditioning runs. To our knowledge, the only report of such behavior of permeate flux during membrane conditioning is the communication by Marinas [45] We attribute the observed flux decrease to the gradual change in Rm as the salt sorbs to the membrane material, which binds more water and swells as a result. It should be noted that pure water flux test can not be used to measure the increased resistance of the membrane to pure water because exposing the membrane to pure water would reverse the electrolyte-induced changes in the membrane structure. More experiments were conducted under strictly laminar flow and Eqs. (2.4) and (2.5) were used to determine the osmotic pressure to factor out the flux decline due to the change in Rm alone. It was found that the correction factor a defined as: a =Rm/R0 (2.12) was in the (1.02 to 1.06) range. The value of a =l.04 was chosen in calculations of the true osmotic pressure. Taking the correction factor into account is especially important for low osmotic pressures. For instance, an osmotic pressure of 20 psi (0.138 MPa) would be overestimated to be 30 psi (0.206 MPa) at a transmembrane pressure differential of 300 psi (2.068 MPa) if the correction factor was not taken into the account. 47 Table 2.1. R0 experimental protocol .UQF'JT".C, Task a Measured ' values 1 Membrane compaction J 0 (duration 60 h) 2 Measurement of the hydraulic J 0 Rmo (2.1) resistance of the membrane to pure water .1. 7‘. ,‘. __..‘.‘ . 7 . . .1 '.' ’1 r-v f» 1.1 Yrs-why ‘ .. . Urqfi.‘ .~o,.l"1114' 1w. ’. - ‘4 v ,1 U‘Hf' . ‘1'“ .- L k! T 9.1' . ‘v - coditioning (duration 12 h) (2.6) 4 Measurement of the hydraulic J , C resistance of the membrane to pure Air," (2.4), (2.5) water after membrane conditioning b Rm (2'7), Rd = 0 5 Measurement of the osmotic pressure J , C p Ayrm (2.7), and salt permeability constant R d = Cm (2.2) _ [R u_ 4303‘ -. in - ‘.',‘ r v . , ‘1 :‘r 11}-‘R- we"! [*1 L:.§‘J$f.’l.gi‘ .:‘F'- r ‘3 . ,3 ,. e“ -' ‘ w" ibut‘i'jiefl - {diff-‘5“; -. ,s.;g,u.. . ' ii ‘. T 1 .T'. ‘ ‘L‘u'..*..{n"f..'f‘.‘y..'_; . . ._ , 6 Mebran fouling experiment I I J C p d . Robs (2.6) (duration ca. 24 h) Aflm (2.10), B measured at step 5 Cm (2.2) Rd (2.7) e (2.11) a . . . . ( Values maintained constants during the experiments are AP , T , and retentate flowrate. Unless mentioned otherwise, the flow regime was transitional with Re-Sc-dh/L=ll,640 this step was performed for several membranes only, for which the average correction factor a (Eq. (2.12)) was computed. This average value was used in all other filtration experiments. The flow regime was laminar with Re = 307 and Re- Sc - d h /L = 4,380 .) 48 2.4.3. Measurement of the membrane salt permeability constant I 90 __ I I I 85 __ .‘ Conditioning A ‘ 80 «a .C 0 N O. E '0 75 ‘~ El .'e 3 Compaction ll 5 |*-—- Decline due to Arrm ‘L 65 . , . , 60 Jw I Decline due to change in Rm ll- - - - 4’ Tl * t t i i -60-24-2 0 2 4 6 81012 Time(h) Figure 2.2. Flux change during conditioning, at a transmembrane pressure difference of 300 psi (2.068 MPa) and an ionic strength of 10‘2 M. 49 Table 2.2. Measurement of the membrane salt permeability constant a __ AP J0 Cp C... A”... PF ” B 3° (psi) (1.761.211) (mmoVL) (moi/L) (psi) - (m/s) (m/s) 251 44.70 0.2904 0.0236 16.23 2.36 1.54107 -7 299 51.50 0.2886 0.0269 18.54 2.69 1.5510 - -7 (1.53i:0.01) ~10 7 351 60.34 0.2877 0.0319 22.04 3.19 1.52-10 -7 400 66.69 0.2967 0.0361 24.93 3.61 1.53-10 (Measurements were conducted at the feed ionic strength of 10'2 M. After complete compaction and conditioning, filtrations at different pressures were conducted for at least 2 h to achieve steady state. Re = 307, Sc = 621, k = 1.4210"5 m/S, d}, = 0.0068 m. u The apparent fluxes (without adjustment based on feed solution temperature), which were used to calculate B , using Eqs. (2.3), (2.5), and (2.10). b Polarization factor defined as PF = Cm /Cf. c The average value. The error range corresponds to 95% confidence interval.) 50 A series of experiments were conducted in a laminar flow to determine the membrane salt permeability constant. Crossflow retentate flow rate set at 1.8 L/min translated into the Reynolds number of 307 and Re- Sc - 6% = 4,3 80 < 104 making Eq. (2.4) applicable. After membrane compaction and conditioning were completed, the transmembrane pressure differential was adjusted to achieve a range of different values of C m. Table 2.2 shows the membrane salt permeability constant (B) calculated using Eq. (2.10). The results indicated that B was approximately constant at different pressures that corresponded to different values of permeate flux and C m. This allowed us to use B values to calculate Air," from Eq. (2.10) during filtration experiments when colloids were present in the feed. 2.4.4. Permeate flux, N aCl rejection, and structure of the deposited layer at different ionic strengths The permeate was recirculated back into the feed tank to keep the feed salt concentration constant and to run longer experiments so that the steady state values for permeate flux, rejection and the mass of deposited colloids could be determined. More importantly, maintaining the constant feed composition enabled modeling of the temporal evolution of flux across the membrane in a particular location (e. g. entrance) along the membrane module. This is in contrast to the experimental design when the permeate is discarded; in such situation, the feed concentration increases with time and the permeate flux observed 51 (a) (b) deionized water x 0.9 , 98 Cf: 0.1" a 08 8 ' g as .5 07 "i 3: “1 0.6 i 3 O 5 924 cf 0.01M z 0.5] ; 90 ~—————4—-—r —-4——a—fi 0.4 o 4 a 1216 2024 Time (h) (C) (d) 0.04., «E12 Amos 1° cit-0.01 u d V I .5 3 Cf 0.1" E ‘l'.’ 0 0.01 a ‘ deionized water 3 2 oi _ .- .-—- —.~ --o.1 go . . . . 1 o 4 s 12 1e 20 24 a o 4 a 12 1e 20 24 Timeihi Time (h) (e) (t) 0.6 '1 8 l 1 090.1141 0.5 . E ’ A 6 1 EM :5 0.3 “c, 41 2 s S 0'2 E 2] E 0.1 I o Lava—4--tamer—“4-- - o 4 o 4 8 12 1e 20 24 0 4 8 12 16 20 24 Time (h) Timth) Figure 2.3. (a) Normalized permeate flux at three ionic strengths;, (b) observed salt rejection, Robs ; (c) salt concentration at membrane surface, Cm; ((1) mass of deposited particles, M d ; (e) effective porosity, e , and (f) hydraulic resistance of deposited colloidal layer, Rd. 52 (No spacers were used in these experiments. Transmembrane pressure differentials: - for deionized water (no salt added): AP = 276 psi (1.903 MPa) — for 10‘2 M salt solution: AP = 290 psi (1.999 MPa) - for 10'1 mol/L salt solution: AP = 400 psi (2.758 MPa); Initial conditions - for deionized water (no salt added): J = 54.70 L/(mzh); - for 10'2 M salt solution: J = 55.78 L/(mzh), C,,, = 0.028 mol/L, C1, = 4.38-10 mol/L, Azrm = 19.4 psi (0.134 MPa), B = 2.41107 m/S; - for 10’1 mol/L salt solution: J = 54.45 L/(mzh), Cm = 0.17 mol/L, C p = 2.07103 mol/L, Arr," = 120 psi (0.827 MPa), B = 1.94107 m/s.) .4 53 at different times of the filtration experiment corresponds to permeate flux at different times and at different locations along the membrane module, which makes flux and rejection data difficult to interpret. Figure 2.3 illustrates transient behavior of J , Robs , Cm , M 5 , and Rd during p , colloidal fouling experiments with 200 mg/L silica suspension at three different ionic strengths. There was no appreciable permeate flux decline for the silica suspension in deionized water (Fig. 2.3a). While it is clear from M p data that particles were depositing on the membrane surface (Fig. 2.3d ), the resistance of the deposit was much smaller than that of the membrane itself and therefore the deposition did not lead to an observable increase in the overall resistance to the permeate flow. Because no flux decline was observed, it was not possible to derive the effective porosity of the particle deposit from the permeate flux data. To estimate the porosity of the deposited layer, another experiment was carried out using polyamide thin-film composite ultrafiltration membrane (Model YMGKSP1905, GE Osmonics). The hydraulic resistance of this ultrafiltration membrane was determined to be 2.84 '1013 m-l. After the full membrane compaction (3 h duration), filtration of 200 mg/L silica particle suspension in deionized water was conducted with the same initial flux of 52.6 L/(m2 h) (1.46 -10'5 m/s) as that employed in R0 experiments. Using Eq. (2.11), the average porosity of the deposit layer was calculated to be 0.63. For non-zero electrolyte concentrations, the permeate flux decreased significantly before a steady state value was reached. The steady state value of the permeate flux was smaller, 54 while the time to reach the steady state was higher, for the higher ionic strength (Fig. 2.3a); the same trend for the steady state flux and the time to reach steady state flux as a function of ionic strength have been reported previously [15, 30]. Based on the permeate flux (Fig. 2.3a) and salt rejection (Fig. 2.3b) data, Cm as a function of filtration time was calculated (Fig. 2.3e, Table 2.1). In the course of the experiment, C m increased to reach a maximum (after 6 h for 10'2 M and after 8 h for 10'1 M ionic strength) and then decreased slightly resulting in an overall increase with respect to the initial value. The initial increase and the resulting overall increase were attributed to the hindrance effect of the fouling layer on salt back diffusion [9]. Possible reasons for the decrease in Cm are discussed in section 2.4.5. Effective porosity a of the particle deposit was calculated to be 0.41 and 0.27 for ionic strengths of 0.01 M and 0.1 M, respectively. This was much lower than that for the case of particles suspended in deionized water. The decrease in porosity was attributed to the lower surface charge and smaller repulsive forces between the particles with increasing ionic strength (section 2.4.1). It should be noted that although more particles were deposited on the membrane surface at 10'2 M than at 10" M (Fig. 2.3d) ionic strength, the deposited layer at 10'1 M had a considerably larger hydraulic resistance (Fig. 2.31) because of the smaller effective porosity (Fig. 2.3e). Both Afl'm and Rd were larger at 10'1 M, which resulted in a larger flux decline at 10'1 M (Fig. 2.3a). Figure 2.3e provides details on the evolution of the effective porosity of the deposit. For both electrolyte concentrations, the effective porosity increased in the first hour and then gradually decreased for the remainder of the experiment. To verify whether the initial 55 increase in porosity corresponded to fractal growth, the following scaling expression [46] was applied: 2 oc N f , where I; is the mean surface height of the deposit, N is the number of particles deposited, and f is the scaling exponent that relates to the fractal dimension D via D =1+l/ f . N was calculated from the mass of the deposit, M p , while il- was estimated from M p and the porosity, 8 using Eq. (2.11). With a very good scaling observed ( p2 > 0.996 ), the fractal dimension was calculated to be 2.1 — the value that is characteristic of reaction-limited conditions of deposit growth. The decrease in deposit porosity in the later stages of filtration was attributed to the processes of breakup and restructuring. Analysis of the permeate flux and rejection data at the very early stages of colloidal fouling can help to elucidate how newly deposited particles interact with the concentration polarization layer of salt. To record such data, a series of R0 experiments with silica suspension of two solid fractions (200 mg/L and 500 mg/L) was conducted in laminar flow. The experimental sequence for each test was: (1) complete membrane compaction (60 h); (2) membrane conditioning (10'2 M, 12 h); (3) introduction of silica stock solution to achieve the desired silica particle loads, (4) colloidal fouling experiment. Transmembrane pressure differential was maintained at 300 psi (2.068 MPa) without any adjustments of system pressure. This ensured that the salt concentration profile was well stabilized before the introduction of particles, and that any changes in permeate flux and rejection were due to the presence of silica particles. As Fig. 2.4 demonstrates, the introduction of colloids resulted in a short term increase in Robs . For 200 mg/L silica 56 """" (a) 100 100 I’ 98 98 W S 96 :5 90 ;‘ I 3 ,i 3 Conditioning ; ,' Filtration O ,' I .. o " l, q a 9‘ : I m 94 '1 'l' 92 Conditioning"; ‘ Flltratlon 92 j; ' E m i l I "l," J , . . 90 |____ _ __T_ 'l'jr_. _. -4 .3 .2 4.11 o 1 2 3 4 .4 .3 .2 £4 0 1 2 3 4 :‘Tlmo (h) .17 Time (h) (c) (d) :' 0 10° ' 100 ' 99 " 99 ‘ 3 1: Q: a; 98 E. ooooooOoOoOOii 9° .0 fl 0 000000000000 .0 T % n: o 97 n: ‘bqb 97 - c°"dm°"'"9 ”hawk Conditioning Flltratlon 0% 96 r T :7. l l ‘1 96 h 7 I r I $91 450 40 -20 o 20 4o 60 .50 .40 40 o 20 4o 60 Time (Will) Time (Will) Fi are 2.4. Effect of introduction of silica particles on initial observed salt rejection at 10' M: (a) 200 mg/L silica particles; (b) 500 mg/L silica particles. (0) and (d) show the initial filtration in detail for (a) and ((1), respectively. 57 concentration, the initial rejection increased from 97.5% to 97.9%; for 500 mg/L silica concentration, the increase was from 98.3% to 99.2%. Similar increase in the initial rejection during nanofiltration (NF) of 200 mg/L colloidal loading, 10'2 M NaCl feed suspension was reported earlier [9]. Because the conductivity measurement interval (2 min for filtration stage) was larger than the conductivity cell hydraulic detention time (0.6 min), the real initial increase in Robs was even larger than what the data suggested. It should be noted that upon addition of the stock silica suspension (pH 9.7 i 0.1) the particles-free feed water (pH 6.1 to 6.80, the pH of the resulting suspension was in the range of 7.1 to 7.9. Because of the known dependence of membrane rejection on pH, rejection of the membrane was recorded as a function of pH. It was found that an increase in feed solution pH from 6.1 to 7.9 resulted in an increase in salt rejection of less than 0.1%. Another control experiment was conducted to ascertain that the observed increase of rejection was due to the added colloids and not due to changes in solution chemistry. First, the stock silica suspension was filtered through a 10 kDa ultrafiltration membrane to get 9.7 mL solution free of particles. The 10 KDa membrane was chosen to ensure 1) the complete removal of silica particles and b) zero rejection of all solutes preset in the silica suspension. Second, membrane compaction and conditioning was conducted using the same approach as that used in all R0 experiments conducted in this work. Permeate conductivity was continuously recorded. Third, after membrane conditioning, the 9.7 mL of the particle-free solution was added into the feed tank. During the experiment no pressure adjustment was made to ensure that there was no measurable change in permeate 58 conductivity upon the addition of silica suspension, which confirmed that the increase in salt rejection was indeed due to the effect of added silica colloids. 2.4.5. Hypothesis of the concentrated flowing layer of silica particles and its depolarizing effect on NaCl transport The observed transient behavior of salt rejection and permeate flux can be explained by taking into the account the complex structure of colloidal deposits on membrane surfaces; such deposits consist of a stagnant layer and a concentrated flowing layer. At the earlier stages of colloidal fouling, the mass of colloidal particles in the flowing layer is significant in comparison with the mass of particles in the growing stagnant layer. Under these conditions, mixing of the salt concentration polarization layer by the concentrated flowing layer of colloidal particles is relatively important with respect to the enhancement of the concentration polarization due to the formation of the stagnant colloidal layer. As the stagnant part of the deposit grows, the relative importance of mixing becomes increasing smaller and the rejection decreases. The analysis of transient behavior of J, Robs , Cm, and M p for 10'2 M ionic strength case shows that after the mass of deposited particle continues to grow while the permeate flux remains stable at the steady state value. It can be concluded from these observations that the depositing particles do not get incorporated into the stagnant deposit. Instead they become a part of the concentrated flowing layer, which results in an increase in the mixing of the salt concentration polarization layer. This proposed scenario is corroborated by the observed gradual decrease in Cm and the corresponding increase in 59 observed rejection afier the steady state permeate flux is achieved. Ng and Elimelech [17] also reported an increased Robs after flux rate had stabilized. At 0.1 M ionic strength, the depolarizing effect of the concentrated flowing layer is less pronounced. With Cm (Fig. 2.3e) and M p (Fig. 2.3d) remaining almost constant afier 8 h of the experiment, the gradual decrease in permeate flux (Fig. 2.3a) appears to be due entirely to the gradual changes in cake porosity (Fig. 2.36) 2.4.6. Evolution of relative contributions of osmotic pressure and resistance of the colloidal deposit to the permeate flux decline To illustrate the relative importance of osmotic pressure and colloidal deposit, we adapt the graphical approach used by Elimelech and co-workers [9, 17] wherein expression for permeate flux (Eq. (2.7)) is re-written as AP = AP," (t) + APd (t) + Azrm (t) , or: 1=(Apm(t)+APd(t)+Aflm(t)], (2.13) APAP AP where AP is the transmembrane pressure differential, AP," = ,ulRm is the pressure drop due to the hydraulic resistance of the membrane, and APd = ,ule is the pressure drop due to the hydraulic resistance of the deposited layer of colloidal particles. Equation (2.13) can be interpreted as an energy balance that describes how the input energy is expended in overcoming the three resistances to the flow (Fig. 2.5). Dashed line in Fig. 60 2.5 corresponds to the osmotic pressure Afl'm in the absence of colloidal fouling. At steady state, the contribution of the colloidal layer resistance, APd , to the flux decline was about the same for both ionic strength and was ca. 18 %. The relative enhancement of the osmotic pressure was also approximately the same in both cases but due to the relatively higher importance of the osmotic pressure at the higher ionic strength, cake-enhanced osmotic pressure contributed significantly to the flux decline only at 10'1 M. For example, while the steady state contributions of membrane resistance and additional resistances related as 70% to 30% in the absence of colloids, the relationship became 42% to 58% with colloidal foulants present in the feed. At 10'2 M, except for the first 3 h of filtration, pressure drop due to colloidal layer was the more significant “energy sink”. Thus, in contrast to the case of filtration of colloids suspended in deionized water, colloidal fouling in the presence of even small amount of electrolyte (such as 10'2 M) becomes significant. The changes in colloid layer porosity and resistance in the presence of electrolyte have to be accounted for to predict permeate flux accurately. For example, when the cake-enhanced concentration polarization model [9, 16] was applied to our experimental data and a constant porosity of 0.4 was assumed, the Cm value calculated from the model was generally more than 3 times higher than those shown in Fig. 2.3c. At the end of the experiment with 10'l M ionic strength, this would produce the osmotic pressure higher than the operating transmembrane pressure differential of 400 psi (2.758 MPa). 2.4.7. Effect of the feed channel spacers on N aCl rejection and permeate flux 61 (a) (b) V o 5 80“ “g- 60“ Apmk 'g ‘5. g 604 §45iAflmo 3 ,0, gum. >| APd% A'I'l'mok >. APd% U: 20- 9151 '5 pal . C C "4 oY T—._—_f-—A-, 11.1 of r , T A A 0 4 81216 2024 0 4 312162024 Time ('1) Time "ll Figure 2.5. Energy consumption analysis for the system: (a) at 10'2 M; (b) at 10'1 M. (operating conditions are listed in the caption for Fig. 2.3.) 62 To shift the balance from the formation of the (undesirable) stagnant layer of deposited particles to the formation of the (desirable) concentrated flowing layer, additional mixing in the vicinity of membrane surface is needed. An accepted method of providing such mixing in membrane systems is the use of membrane spacers. This goal as well as the need to test our hypothesis in the context of industrial applications motivated the part of this study where feed channel spacers were used. Figure 2.6 shows the normalized flux and observed salt rejection as a function of filtration time when a spacer was placed in the feed channel. Filtration of silica suspension in deionized water was not conducted because there was no flux decline observed even when spacers were not used (Fig. 2.3a). When the spacers were used, no flux decline was observed at 10'2 M, and the flux decline was much smaller at 10'1 M compared with runs without the spacers (Fig. 2.3a). lmportantly, after the initial increase caused by the introduction of particles into the feed, Robs remained stable at this elevated level. This was despite a decrease of particle concentration in the feed as detected by UV absorbance. It appears that the effect of spacers is that of “trapping” of particles in the interfilament domain of feed channel spacers where particles remain without depositing on the membrane surface. In terms of the concentrated flowing layer model, introducing a spacer inhibits the formation of the stagnant deposit so that mixing by colloidal particles in the flowing layer becomes the dominating effect of particles on the salt concentration polarization. 63 (a) 1 °- 5 0'9“ c 001ME "" f= . E 0.8- / g 01‘ Cf= 0.1M N E 0.6- 0 2 0.5- 0.4 '“_“‘—“”T—"‘“——i———~ —r— —-—~—T —~T—-——— 1 o 4 8 12 16 20 24 Time (h) (b) i Cf=0.01M Cf=0.1M g 96 0 m 95 Conditioning Filtration °‘ 1. 93 r—— “—4 . . . A .4 o 4 812162024 Time (it) Figure 2.6. Effect of the feed channel spacers on membrane performance ((a) Normalized permeate flux; (b) Observed salt rejection. Silica particle loading: 200 mg/L; Operating transmembrane pressure differential: for 10'2 M, AP = 250 psi (1.723 MPa), and for 10'1 M, AP = 330 psi (2.274 MPa); Initial flux: for 10'2 M, J = 55.7 L/(mzh) (1 5410'5 m/s) and for 10‘1 M J = 60.0 L/(mzh) (1.55.10'5 m/s).) 64 Large “fluidized particles” have been successfillly applied in ultrafiltration with tubular membranes: mass transfer was shown to improve when 0.7 mm and smaller glass particles were used as turbulence promoters in ultrafiltration of polyethylene glycol [47] and when 3 mm stainless steel particles were used in the separation of gelatin solutions [48]. As our results indicate, colloid-sized particles can be used to achieve improvements in flux and rejection for salt rejecting membranes as well. This opens new, perhaps counter-intuitive, possibilities for controlling the salt rejection and permeate flux by using spacers in conjunction with introducing particles with low deposition propensity into the feed. 2.5. CONCLUSIONS Individual contributions of osmotic pressure and colloidal fouling to the RO permeate flux decline can be identified by determining salt permeability constant and measuring salt transport across the membrane. By using this approach, we demonstrate that the two- way coupling between salt concentration polarization and colloidal deposition is essential in determining both permeate flux and rejection. On one hand, porosity of the deposit and its resistance to the permeate flux are measured to be strong functions of the solution ionic strength. The initial growth of the colloidal deposit on the surface of the salt- rejecting membrane is found to be fractal with the fractal dimension of 2.1 that corresponds to reaction-limited aggregation. After the breakup and restructuring of the deposit, its resistance increases dramatically and contributes significantly to the overall flux decline. On the other hand, deposition of colloidal particles on the membrane surface 65 influences salt transport to and across the membrane in a complex way. Formation of the stagnant colloidal deposit results in the hindrance of the back diffusion of salt away from the membrane surface. Under certain conditions, however, the presence of colloidal particles improves membrane performance. Specifically, the introduction of colloidal particles into the feed results in a short term increase in salt rejection; this increase can be sustained over the long term when feed channel spacers are used. Based on the above observations, we hypothesize that the concentrated flowing layer of colloidal particles is responsible for the local mixing of the concentration polarization layer of salt and that the overall effect of colloidal deposition on permeate flux and rejection is determined by the balance between the stagnant and flowing parts of the colloidal deposit. These findings point to the potential of using particles with low deposition propensity as “mobile mixers” to complement feed channel spacers as means of improving performance of salt- rejecting membranes. ACKNOWLEDGEMENTS Financial support of this work by the National Water Research Institute (project no. 05- TM-007) is gratefully acknowledged. We also thank Dow-FilmTec for providing us with membrane samples and Nissan Chemical America Corp. for supplying the silica suspension. 66 NOMENCLATURE (WOW: 5 h membrane filtration area (m2) solute permeability constant (ms-1) solute concentration at membrane surface (molL'l) solute concentration in permeate (molL'l) feed solute concentration (molL'l) solute diffusion coefficient in water (mzs'l) hydraulic diameter (m) particle diameter (m) channel height (m) volumetric permeate flux (ms'l) volumetric pure water permeate flux (ms'l) salt flux (molm'zh'l) channel length (m) mass of the particles deposited on membrane surface (kg) transmembrane pressure differential (MPa) pressure drop due to resistance of the membrane (MPa) pressure drop due to resistance of the deposited colloidal layer (MPa) universal gas constant (J K'lmol'l) resistance of the deposited colloidal layer (m'l) initial hydraulic resistance of the membrane to pure water (m'l) hydraulic resistance of the membrane to pure water after membrane conditioning (m'l) observed solute rejection absolute temperature (K) correction factor osmotic pressure across the membrane (MPa) dynamic viscosity of water (kgm'ls'l) concentration polarization boundary layer thickness (m) effective porosity of the deposited colloidal layer particle density (kgm'3) correlation coefficient 67 REFERENCE 10. 11. 12. 13. 14. M. Barger and RP. Carnahan, Fouling prediction in reverse osmosis processes, Desalination 83 (1991) 3-33. Committee Report: Recent advances and research needs in membrane fouling, J. Am. Water Works Assoc. 97 (2005) 79-89. M. Luo and Z. Wang, Complex fouling and cleaning in-place of a reverse osmosis desalination system, Desalination 141 (2001) 15-22. K. T. Chua, M. N. A. Hawlader and A. Malek, Pretreatment of seawater: Results of pilot trials in Singapore, Desalination 159 (2003) 225-243. Y. R. Ning, L. T. Troyer and S. R. Tominello, Chemical control of colloidal fouling of reverse osmosis systems, Desalination 172 (2005) 1-6. B. G. Darton and M. Fazel. A statistical review of 150 membrane autopsies. in International Water Conference. 2001. J. W. McCutchan and J. S. Johnson, Reverse osmosis at Coalinga, California, J. Am. Water Works Assoc. 62 (1970) 346. S. 1. Graham, R. L. Reitz and C. E. Hickman, Improving reverse osmosis performance through periodic cleaning, Desalination 74 (1989) 113-124. E. M. Hoek and M. Elimelech, Cake-enhanced concentration polarization: a new fouling mechanism for salt-rejecting membranes, Environ. Sci. Technol. 17 (2003) 5581-5588. J. M. Jackson and D. Landolt, About the mechanism of formation of iron hydroxide layers on reverse osmosis membranes, Desalination 15 (1973) 361-378. E. Staude and W. Assenmacher, A contribution to fouling in hyperfiltration induced by chemical reaction, Desalination 49 (1984) 215-228. R. D. Cohen and R. F. Probstein, Colloidal fouling of reverse osmosis membranes, J. Colloid Interface Sci. 114 (1986) 194-207. S. G. Yiantsios and A. J. Karabelas, The effect of colloid stability on membrane fouling, Desalination 118 (1998) 143-152. J. C. Schippers, J. H. Hanemaayer, C. A. Smolders and A. Kostense, Predicting flux decline of reverse osmosis membranes, Desalination 38 (1981) 339-348. 68 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. X. Zhu and M. Elimelech, Colloidal fouling of reverse osmosis membranes: measurements and fouling mechanisms, Environ. Sci. Technol. 31 (1997) 3654- 3662. E. M. Hoek, A. S. Kim and M. Elimelech, Influence of crossflow membrane filter geometry on colloidal fouling of reverse osmosis membranes, Environ. Eng. Sci. 19 (2002) 357-372. H. Y. Ng and M. Elimelech, Influence of colloidal fouling on rejection of trace organics contaminants by reverse osmosis, J. Membr. Sci. 244 (2004) 215-226. P. Lipp, R. Gimbel and F. H. Frimmel, Parameters influencing the rejection properties of FT30 membranes, J. Membr. Sci. 95 (1994) 185-197. S. Lee, J. Cho and M. Elimelech, Influence of colloidal fouling and feed water recovery on salt rejection of R0 and NF membranes, Desalination 160 (2004) 1- 12. C. W. van Oers, M. A. G. Vorstman and P. J. A. M. Kerkhof, Solute rejection in the presence of a deposited layer during ultrafiltration, J. Membr. Sci. 107 (1995) 173-192. S. Lee, J. Cho and M. Elimelech, Combined influence of natural organic matter (N OM) and colloidal particles on nanofiltration membrane fouling, J. Membr. Sci. 262 (2005) 27-41. A. I. Schafer, A. G. Fane and T. D. Waite, Fouling effects on rejection in the membrane filtration of natural waters, Desalination 131 (2000) 215-224. R. Jiraratananon, D. Uttapap and P. Sampranpiboon, Crossflow microfiltration of a colloidal suspension with the presence of macromolecules, J. Membr. Sci. 140 (1998) 57-66. S. S. Madaeni and A. G. F ane, Microfiltration of very dilute colloidal mixtures, J. Membr. Sci. 113 (1996) 301-312. V. T. Kuberkar and R. H. Davis, Modeling of fouling reduction by secondary membranes, J. Membr. Sci. 168 (2000) 243-258. R. M. McDonogh, C. J. D. Fell and A. G Fane, Surface charge and permeability in the ultrafiltration of non-flocculating colloids, J. Membr. Sci. 21 (1984) 285- 294. K.-J. Hwang, H.-C. Liu and W.-M. Lu, Local properties of cake in cross-flow microfiltration of submicron particles, J. Membr. Sci. 138 (1998) 181-192. 69 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. Y. Lee and M. M. Clark, Modeling of flux decline during crossflow ultrafiltration of colloidal suspensions, J. Membr. Sci. 149 (1998) 181-202. P. Harmant and P. Aimar, Coagulation of colloids in a boundary layer during cross-flow filtration, Colloids Surf, A 138 (1998) 217-230. V. V. Tarabara, I. Koyuncu and M. R. Wiesner, Effect of hydrodynamics and solution chemistry on permeate flux in crossflow filtration: Direct experimental observation of filter cake cross sections, J. Membr. Sci. 241 (2004) 65-78. J. C. Chen, M. Elimelech and A. S. Kim, Monte Carlo simulation of colloidal membrane filtration: Model development with application to characterization of colloid phase transition, J. Membr. Sci. 255 (2005) 291-305. S. Sethi and M. R. Wiesner, Modeling of transient permeate flux in cross-flow membrane filtration incorporating multiple particle transport mechanisms, J. Membr. Sci. 136 (1997) 191-205. J. Schwinge, P. R. Neal, D. E. Wiley, D. F. Fletcher and A. G. Fane, Spiral wound modules and spacers. Review and analysis, J. Membr. Sci. 242 (2004) 129-153. M. R. Wiesner, Morphology of particle deposits, J. Environ. Eng. ASCE 125 (1999)1124-1132. K. J. Hwang and C. L. Hsueh, Cross-flow microfiltration of aggregated submicron particles, J. Membr. Sci. 201 (2002) 137-148. E. F. Leonard and C. S. Vassilief, The deposition of rejected matter in membrane separation processes, Chem. Eng. Comm. 30 (1984) 209-217. R. H. Davis and S. A. Birdsell, Hydrodynamic model and experiments for cross- flow microfiltration, Chem. Eng. Sci. 49 (1987) 3203-3209. D. Airey, S. Yao, J. Wu, V. Chen, A. G. Fane and J. M. Pope, An investigation of concentration polarization phenomena in membrane filtration of colloidal silica suspensions by NMR micro-imaging, J. Membr. Sci. 145 (1998) 145-158. V. Chen, A. G. Fane, S. Madaeni and I. G. Wenten, Particle deposition during membrane filtration of colloids: transition between concentration polarization and cake formation, J. Membr. Sci. 125 (1997) 109-122. M. Mulder, Basic Principles of Membrane Technology. 2nd ed. 1996, Dordrecht: Kluwer Academic Publishers. R. Rautenbach and R. Albrecht, Membrane Processes. 1989, New York: John Wiley & Sons Ltd. 70 42. 43. 45. 46. 47. 48. H. K. Lonsdale, U. Merten and R. L. Riley, Transport properties of cellulose acetate osmotic membranes, J. Appl. Polym. Sci. 9 (1965) 1341-1362. P. C. Carman, Fluid flow through a granular bed, Trans. Inst. Chem. Eng. London 15 (1937) 150-166. W. J. Koros, Y. H. Ma and T. Shimidzu, Terminology for membranes and membrane processes, Pure & Appl. Chem. 68 (1996) 1479-1489. B. J. Marinas, Effects of solute concentration and membrane compaction on the performance of a fully aromatic polyamide reverse osmosis membrane, in National meeting - American Chemical Society, Division of Environmental Chemistry. 1996. p. 125-126. S. Veerapaneni and M. R. Wiesner, Deposit morphology and headloss development in packed bed filters, Environ. Sci. Technol. 31 (1997) 2738-2744. M. J. van der Waal, P. M. van der Velden, J. Koning, C. A. Smolders and W. P. M. van Swaay, Use of fluidized beds as turbulence promoters in tubular membrane systems, Desalination 22 (1977) 465-483. G. M. Rios, H. Rakatotoarisoa and B. Tarodo de la Fuente, Basic transport mechanics of ultrafiltration in the presence of fluidized particles, J. Membr. Sci. 34 (1987) 331-343. 71 S.2. SUPPLEMENTARY DATA 8.2.1. Determination of the salt permeability As demonstrated in Section 2.4.3, A series of experiments were conducted in a laminar flow to determine the membrane salt permeability constant with varied feed concentration. Table 82.1 shows the membrane salt permeability constant (B) calculated using Eq. (2.10). The results indicated that B was approximately constant at different pressures that corresponded to different values of permeate flux and C m . Table 82.1. Measurement of the membrane salt permeability constant Fee concentrgtion AP J0 Cp Cm A” m PF B (Cf ) (psi) (1&1211» (mmol/L) (mol/L) (psi) - (m/s) 251 44.70 0.2904 0.0236 16.23 2.36 1.54107 0.001 M 299 51.50 0.2886 0.0269 18.54 2.69 155-10': 351 60.34 0.2877 0.0319 22.04 3.19 1.5210 400 66.69 0.2967 0.0361 24.93 3.61 1.53107 250 38.76 0.5721 0.9809 38.97 2.11 0978-107 0003 M 302 45.66 0.5514 0.9816 51.50 2.42 0969-10“: 362 53.96 0.5730 0.9809 53.64 2.84 1013-10 400 58.95 0.5505 0.9817 67.78 3.13 0963-107 250 34.10 0.8871 0.0966 66.71 1.93 0.876107 0005 M 304 41.99 0.8295 0.1126 77.93 2.25 0863-10“: 353 47.43 0.8223 0.1252 86.69 2.50 0869-10 402 53.69 0.7908 0.1415 98.05 2.83 0836-107 250 31.46 1.0689 0.1286 88.9 1.84 0.73010“7 0007 M 310 38.81 0.9789 0.1484 102.8 2.12 0714-10“: 352 45.22 0.9807 0.1681 116.5 2.40 0735-10 399 49.57 0.9987 0.1830 126.8 2.61 0754-107 247 27.15 1.3929 0.1521 105.1 1.69 0.695-10’C 0.009 M 302 32.82 1.3227 0.1698 117.5 1.89 0714-10: 350 38.45 1.3659 0.1894 131.1 2.10 0774-10 400 46.60 1.2966 0.2220 153.9 2.47 0758-107 250 26.82 1.9509 0.1845 127.3 1.68 0794-107 0.0“ M 300 32.81 1.7709 0.2074 143.3 1.89 0783-10: 354 40.08 1.7799 0.2388 165.2 2.17 0.834-10 401 45.65 1.7259 0.2662 184.3 2.42 0.825107 72 8.2.2. Imaging the fouling layer using transmission electron microscopy (TEM) Fig. 82.1 shows a cross-section image of the deposit of the silica particles formed at 0.1 M using transmission electron microscopy (TEM). The cross-section was imaged using Joel 6400V scanning electron microscope (Japan Electron Optics Laboratories, Japan) with a field emission (Oxford EDS, Oxford Instruments, UK) in backseattered electron (BSE) imaging mode with an accelerating voltage of 20 kV. We used following procedures to prepare the cross-section samples of fouled membranes for imaging. Samples were infiltrated with resin of Poly/Bed Mixture (Polysciences, Warrington, PA) in gelatine capsules in agitation for 24 h. Samples were polymerized at 60 °C for 24 h and were cross-sectioned with a thickness of 80-100 nm using an ultramicrotome (Model: MTX, RMC, Boekeler Instruments, Tucson, AZ) equipped with a diamond knife. Fig. 82.1. TEM imaging of the deposit layer formed at 0.1 M 73 CHAPTER THREE Wang, Fulin; Tarabara, Volodymyr V. (2008). On the identification of pore blocking mechanisms during early stages of membrane fouling by colloids. (Submitted for publication in Journal of Colloid and Interface Science) 74 CHAPTER THREE ON THE IDENTIFICATION OF PORE BLOCKING MECHANISMS DURING EARLY STAGES OF MEMBRANE FOULING BY COLLOIDS ABSTRACT A method based on a simple linear regression fitting was proposed and used to determine the type, the chronological sequence, and the relative importance of individual fouling mechanisms in the experiments on the dead-end filtration of colloidal suspensions with membranes ranging from loose ultrafiltration (U F) to non-porous reverse osmosis (RO). For all membranes, flux decline was consistent with one or more pore blocking mechanisms during the earlier stages and with the cake filtration mechanism during the later stages of filtration. For ultrafiltration membranes, pore blocking was identified as the largest contributor to the observed flux decline. The chronological sequence of blocking mechanisms was interpreted to depend on the size distribution and surface density of membrane pores. For salt-rejecting membranes, the flux decline during the earlier stages of filtration was attributed to either intermediate blocking of relatively more permeable areas of the membrane skin, or to the cake filtration in its early transient stages, or a combination of these two mechanisms. The findings emphasize the practical importance of the clear identification of the mechanisms of colloid-membrane interactions as determining the potential for the irreversible fouling and the efficiency of membrane cleaning. 75 3.1. INTRODUCTION Blocking filtration laws describe the following four mechanisms of membrane fouling by colloidal particles (Fig. 3.1): complete pore blocking (also called pore sealing), standard pore blocking (also called pore constriction), intermediate pore blocking, and cake filtration. For the complete blocking, it is assumed that each particle reaching the membrane blocks a pore without superimposing over other particles. For the standard blocking, it is assumed that particles deposit within pores and the pore volume decreases proportionally to the volume of deposited particles. For the cake filtration, it is assumed that depositing particles do not block pores either because membrane is dense and there are no pores to block or because the pores are already covered by other particles and therefore are not available to block. For the intermediate blocking, it is assumed that some particles deposit on other particles (as in cake filtration) while other particles block membrane pores (as in complete blocking). Blocking filtration laws were first proposed by Hermans and Bredée [1] and subsequently developed by Gonsalves [2]. In an experimental study with several membranes, Grace [3] determined that standard blocking occurred with each microfilter examined. Hermia offered a unified analytical description of all four blocking mechanisms and extended their application to power-law non-Newtonian fluids [4]. Although initially developed for a membrane with parallel cylindrical pores, blocking laws have been shown to have predictive power also in applications involving polymeric 76 ill . Figure 3.1. Schematic illustration of the four fouling mechanisms: (a) complete blocking, (b) standard blocking, (c) intermediate blocking, and ((1) cake filtration. ) (C) (d) 77 membranes with irregular pore morphology [5-7]. The blocking laws have been used for studying fouling of porous membranes filtering suspensions of microorganisms [8], proteins [9-11], natural organic matter (NOM) [12], as well as natural water [6] and wastewater [13]. Blocking laws have been used to interpret permeate flux behavior for non-porous membranes as well [14, 15]. However, to our knowledge there were no studies with salt-rejecting (NF, RO) membranes where the effect of the change in osmotic pressure during filtration was duly considered and the separation was carried out in the dead-end geometry. It should be noted in this regard that the theory of blocking laws assumes no back-transport of the colloids from the vicinity of the membrane surface to the bulk of the feed suspension; this assumption narrows the applicability of blocking laws to unstirred dead-end filtration [4], [16]. For membrane filtration carried out in a constant transmembrane pressure mode with the feed flow normal to the membrane surface and with foulants of spherical shape that are completely retained, equations describing the relationship between the total filtered volume, V , and filtration time, t, for each of the four fouling mechanisms are as follows [1,4]: K bV = Q0 (1 _ e-be) (complete blocking), (3.1) K St _ t 1 (standard blocking), (3,2) 2 H V Q0 78 Ki V = ln(l + K tht) (intermediate blocking), (3-3) KCV = 3t _ _2_ (cake filtratlon), (3.4) VQ0 where Qo is the initial flow rate, and K is the constant with the subscript indicating the blocking mechanism. The first three mechanisms are categorized as pore blocking. Hermia [4] demonstrated that a common characteristic equation for different fouling mechanisms can be derived by differentiating Eqs. (3.1-3.4): 2 n _d___t_ = k .‘1’. , (3,5) dyz dV where k is a constant and n is the blocking index equal to 2, 1.5, 1, or 0 for complete blocking, standard blocking, intermediate blocking, and cake filtration, respectively. 2 Equation (3.5) has been used to identify blocking mechanisms by plotting —d-—;— against dV it— and determining the value of the blocking index n [9, ll, 12, 17-19]. The method is d V advantageous in that based on a single plot one can, in principle, i) identify all blocking mechanisms and ii) determine whether the data can be interpreted in terms of blocking laws; for example, in several studies the value of the blocking index n for later stages of filtration was reported to be negative [9, l 1, 12, 17, 18] indicating the failure of blocking laws to explain the flux data. A significant disadvantage of the direct application of Eq. 79 (3.5) is the high sensitivity of the value of n with respect to the noise in the flux data and, relatedly, to the choice of the time interval used in computing derivatives in Eq. (3.5). An alternative approach to identifying mechanisms of colloidal fouling is the application of the blocking laws in their integrated forms, Eqs. (3.1-3.4). Grace [3] identified standard blocking (Eq. (2)) by observing a linear dependence when the experimental data were plotted as t versus ~19; for the early stages of filtration, the data did not fall on a straight line implying that fouling mechanism(s) other than standard blocking led to the initial flux decline. Similarly, cake filtration (Eq. (3.4)) was identified as the dominant fouling mechanism by observing the linear relationship between V and Tt; [19-21]; in these studies, the data corresponding to the earlier stages of filtration did not follow the linear dependence, revealing that some form of pore blocking occurred prior to the start of the cake filtration regime. The linearity of the integrated expressions for standard blocking (Eq. (3.2)) and cake filtration (Eq. (3.4)) makes their use straightforward and facilitates interpretation of the results. Expressions for complete blocking and intermediate blocking given by Eqs. (3.1) and (3.3), respectively, are not linear but linear expressions can be derived by differentiating experimental V(t) data to obtain permeate flux, Q: Q = Q0 - KbV (complete blocking) (3.6) 1 1 . . . — = — + K1" (1ntermedlate blocking) (3.7) Q Q0 80 The main advantage of the use of linear expressions (Eqs. (3.2), (3.4), (3.6), and (3.7)) is that it does not require additional data processing; instead, it involves only a straightforward linear least squares fit, with an easily determinable correlation coefficient, and allows for a facile identification of individual fouling mechanisms. A clear identification of and differentiation between the blocking mechanisms has three important practical implications in that the knowledge of fouling mechanisms can guide decisions on the optimal choice of the membrane and membrane cleaning strategy [22]. It should be noted that although colloidal deposition during the early stages of filtration is the most likely cause of irreversible fouling [22], little is known about the early stages of colloidal fouling of porous (with irregular pore morphology) and, especially, non-porous membranes. The need in such information motivated this study. In this work, we conducted unstirred dead-end filtration experiments with membranes spanning a range of molecular weight cut-offs. We then fit Eqs. (3.2), (3.4), (3.6), and (3.7) to the experimental flux data using the linear least square fitting to identify i) the type, ii) the chronological sequence, and iii) the relative importance of individual fouling mechanisms. 3.2. Materials and methods 3.2.1. Membranes and colloids Table 3.1 lists the membranes used in this study. Prior to filtration experiments, each membrane was soaked in deionized water for 24 h at 4 °C with water exchanged after the first 12 h of storage. Silica colloids (SnowTex-XL, Nissan Chemical America Corp., 81 Houston, TX) were used as the model colloids. Using dynamic light scattering (ZetaPALS, Brookhaven Instrument Corp., Holtsville, NY) the mean hydrodynamic diameter of colloids was determined to be (55.6 i 1.3) nm. 3.2.2. Filtration experiments Dead-end filtration experiments were conducted at a constant transmembrane pressure using a stainless steel filtration cell (HP4750, Sterlitech Corp., Kent, WA) without stirring. The pressure was delivered from a tank of compressed ultrapure zero grade air. The permeate flux was recorded continuously using an electronic mass balance (AV8100, Ohaus Corp., Pine Brook, NJ) interfaced with a computer. For each experiment, 1 L of deionized water was filtered to compact the membrane first; then, clean water permeate flux was recorded for a series of transmembrane pressure values to determine the hydraulic resistance of the compacted membrane; finally, a 200 mL of 200 mg(SiOz)/L aqueous suspension of silica colloids was poured into the cell and the membrane fouling experiment was started. Permeate flux was calculated based on the mass of permeate collected on the balance. The transmembrane pressure was adjusted to bring the initial fluxes in experiments with different membranes as close one to another as possible; however, because of the very wide difference in permeability of different membranes, it was not possible to set the initial flux at the same value. The initial fluxes were 23.4 urn/s, 26.9 urn/s, 108.9 urn/s, 133.7 urn/s, and 189.3 um/s for BW30-365, NF90, UF-30kDa, UF- lOOkDa, and UF-300kDa membranes, respectively. 82 Table 3.1. Membranes used in this study Membrane Membrane Nominal Hydraulic Manufacturer Model number type material MWCO resistarlce , RO PA‘ 100 to 200 Da 1.1-10l4 Dow-FilmTech BW30-365 NF PA 200 to 400 Da 3.8-10l3 Dow-FilmTech NF90 UF PVDFb 30 kDa 1.8-1012 GE Osmonics YMJWSP3001 UF PANc 100 kDa 121012 GE Osmonics YMMWSP1905 UF PESd 300 kDa 4.8-10” Millipore Corp. PBMK a Polyamide; b Polyvinylidene fluoride; c Polyacrylonitrile; d Polyethersulfone 83 3.2.3. Scanning electron microscopy (SEM) analysis Small coupons were cut from the membranes and were soaked in deionized water at 4°C overnight. After the coupons were dried in air, they were first coated for 20 s using a pure osmium coater (N eoc-AN, Meiwa Shoji Co. Ltd., Kyoto, Japan), and then imaged using Hitachi S-47OOII Field Emission scanning electron microscope (Hitachi High Technologies America, Inc. Pleasanton, CA) with a secondary electron detector. The microscope was operated in ultra-high resolution mode with an accelerating voltage of 15 kV and an emission current of 10 11A. 3.2.4. Data fitting procedure The following procedure was adapted to assign portions of the permeate flux dataset to a particular fouling mechanism: i) a portion (“core”) of the entire filtration dataset that fit one of the linear Eqs. (3.2), (3.4), (3.6), and (3.7) with a coefficient of determination higher than 0.99 was identified; ii) the “core” was expanded data point by data point by evaluating if a data point immediately adjacent to the “core” belonged to the 95% confidence interval computed for this data point based on the linear fit for the “core” using a t-test [23]. If the data point was within the interval, the data point was included into the “core” and a new fitting line was determined. Then, the t-test was repeated for the next adjacent data point. The process was stopped when the next data point was found to be located outside of the interval. Each filtration experiment was repeated once and very reproducible fitting results were obtained. 84 3.2.5. Determination of the effective porosity of the colloidal deposit The hydraulic resistance, Rd , of a layer of spherical incompressible uniform particles can be estimated using Kozeny-Carman equation: 1801— M Rd: (2‘? A”, (3.8) ppdpe where A is the membrane filtration area, a is the effective porosity of the particle deposit, M p is the mass of particles deposited on membrane surface , and tip and p p are the diameter and the density of the particles, respectively. M p is estimated based on the feed concentration of silica colloids and the permeate volume. This approach presumes that the back transport of silica colloids from the membrane surface to the bulk of the feed is negligible, which is consistent with the assumptions of the blocking laws theory. When filtration experiments are conducted at a constant transmembrane pressure, the instantaneous value of the hydraulic resistance of the cake can be determined by: Rd = [J°J_ J]Rma (3.9) where Rm is the hydraulic resistance of clean membrane, J 0 and J are the initial and instantaneous values of the permeate flux, respectively. Combining Eqs. (3.8) and (3.9), the effective porosity, a , can be determined. 85 3.3. Results and discussion 3.3.1. Flux performance of different membranes Figure 3.2a shows the normalized permeate flux as a function of permeate volume in experiments with different membranes. The larger decline in the normalized permeate flux observed for the ultrafiltration membranes was partly due to their lower hydraulic resistance (measured in a clean water flux test prior to the fouling experiments); for lower resistance membranes the additional resistance to the permeate flow due to the fouling layer has a relatively higher impact on the flux. Another factor responsible for the membrane-to-membrane difference in the transient behavior of permeate flux was the difference in the type and timing of fouling mechanisms. In Fig. 3.2b the results of filtration experiments for all membranes are presented on the plot of t/V versus V , which should be linear when the fouling mechanism is cake filtration (Eq. (3.4)). While the cake filtration appears to be the primary fouling mechanism for all membranes during the later stage of filtration, the non-linearity of the curves during the early stages of experiments implies that pore blocking of one or more types preceded the cake filtration stage. Similar trends were reported by others [19-21] but fouling mechanisms acting during the early stage of filtration were not clearly identified. The following discussion focuses on this early stage of filtration experiments. Figure 3.3 illustrates the chronological sequence of fouling mechanisms occurring in filtration experiments with different membranes. The error bars correspond to the 95% 86 confidence interval for the beginning and the end of the portion of each experiment where the given blocking mechanism dominates. Although only one mechanism is assigned for one specific section of filtration at the initial stage, it is reasonable to expect that more than one of these blocking processes occurs simultaneously. The blocking mechanism identified in this study as acting during a particular stage of filtration is the one which dominates the fouling process during this stage. During certain periods of filtration, the data could satisfactorily fit more than one linear equation. Such “overlap” has a straightforward and intuitive interpretation as corresponding to the transitional stage wherein two fouling mechanisms act simultaneously. 87 Normalized permeate flux 0.2 0.4 0.3 W (minlmL) 0.2 0.1 0.8 0.6 0.4 ....r\r.(¢...r.n\.;pzrr 1' \ .111 gtfiméfg .kw‘ik El BW30-365 O NF90 <> UF-30kDa D U F-100kDa UF-300kDa l A 1 l l l I r i M :1 BW30-365 i . o NF90 F1 o UF-30kDa _ o UF-100kDa ; i :1 UF-300kDa ’ II .190 A,,,.'.'.'1'.'2. E 1 1 1 1 1 1 l 1 1 1 1 o 50 100 150 200 Permeate volume (mL) Figure 3.2. Normalized instantaneous permeate flux (a) and revere cumulative flux (b) as a function of permeate volume for different membranes. The transmembrane pressures for the BW30-365, NF 90, UF30kDa, UFlOOkDa, UF 300kDa membranes are 300 psi (2.07 MPa), 100 psi (0.69 MPa), 40 psi (0.28 MPa), 30 psi (0.21 MPa), 20 psi (0.14 MPa), respectively. Note: the higher scatter in the flux data at the later stages (Fig. 3a) of experiments with R0 and NF membranes was due to the dissolution of air in the feed resulting in bubbles on the low pressure side of the membrane. 88 3.3.2. Blocking mechanisms for UF membranes In filtration with the UP membranes, Eq. (3.7), but not Eq. (3.6), fitted the flux data well, indicating that intermediate blocking, and not complete blocking, occurred. An example of the fitting approach is shown in Fig. 81 in the Supplementary Data (SD). With UF- 300kDa membrane, standard blocking occurred first, followed by intermediate blocking and then cake filtration (Fig. 3.3). This was due to the large pore size and surface porosity of the UF-300kDa membrane (Fig. 3.4). Particles first deposited in the membrane pores and reduced the area for water to pass through. As expected, it resulted in the fastest flux decline (Fig. 3.2). In contrast, intermediate blocking, standard blocking, and then cake filtration were found to occur successively in filtration experiments with UF-30kDa and UF-lOOkDa membranes (Fig. 3.3). Given the presence of large (up to 500 nm) surface pores in these membranes, standard blocking was clearly occurring at the beginning of the filtration; however, the analysis of the flux data (Fig. 3.3) indicates that intermediate blocking was dominating the overall fouling process. This can be rationalized by accounting for the relatively lower (with respect to UF-300kDa) apparent surface porosity and wider pore size distribution of UF-30kDa and UF-lOOkDa (Fig. 3.4): for such pore morphologies it can be expected that more particles deposit first onto smaller pores and other particles, while relatively fewer particles deposit into the inner channels of larger pores (Fig. 3.4). An explanation based on differences in membrane pore morphology was also offered by Bowen et al. [17] to explain their finding that complete blocking was followed by a standard blocking, and then again by complete blocking: 89 200' '1 ’2 . 100% 2’3 I’3 x/ 1601 a a ,3 111. 5 "l Cake ¢ 2 g $ :5 liitntion é - '120I 5. 5 7 ~/ 60963 i :4 v: 4 a g - .5 V5? Standard ¢ '= 2 I fit blocking ¢ ‘6 so a /_a . / 40% i ,4 at a 4 2' a a E -- -/ intermediate é a. ‘0 $4 é blocking - '/ 20% g I; I é o u a o - o o o 1 o awso- NF- UF- ur- UF- awao- NF- ur- UF- us- 365 90 30110. 10010: 30011011 365 90 soitoa 1ooitoa 300kDa Figure 3.3. The distributions and contributions of the blocking mechanisms in terms of permeate volume (a) and percentage of flux decline (b). 90 such sequence of fouling mechanisms was interpreted to be due to the wide size distribution of membrane pores and foulants. The relative importance of different fouling mechanisms can be evaluated by comparatively analyzing the volume of filtrate collected (Fig. 3.3a) and the percent of flux decline observed (Fig. 3.3b) during the filtration intervals corresponding to individual fouling mechanisms. For UF membranes, the contribution of pore blocking to the overall flux decline was significantly larger than that of cake formation and growth. For all three UF membranes tested, cake filtration corresponded to more than 50% of permeate collected but to less than 25% of the decrease in flux (Fig. 3.3). It was pore blocking that was the main contributor to the flux decline. In experiments with UF- 300KDa membrane for example, standard blocking was the dominant fouling mechanism during the filtration interval wherein only ca. 40% of permeate was collected, but resulted in more than 85% of flux decline. The above comparison is conservative: had the experiments lasted longer, the disparity of the contributions of cake filtration and pore blocking to the overall flux decline would have been even more pronounced. 3.3.3. Blocking mechanisms for salt-rejecting membranes As in the case with UF membranes, cake filtration was clearly identified as the dominant fouling mechanisms responsible for the flux decline during later stages of separation with BW30-365 and NF90 membranes (Fig. 3.3b) as indicated by the very good fit of the experimental data to Eq. (3.4) (Fig. 83.2 in the SD). The part of R0 and NF flux data that corresponds to the collection of the first 45 mL of the permeate could be fit by Eqs. (3.6) 91 and (3.7) equally well; this is consistent with the fact that Eqs. (3.1) and (3.3) have the same filnctional form when linearized to describe the early stages of filtration (Fig. 83.3 in the SD). Thus, it was not possible to differentiate between complete and intermediate blocking mechanisms for R0 and NF membranes based on the fitting results alone. A simple calculation shows, however, that at least 70 layers of silica colloids were deposited on the membrane surface after 45 mL of the feed suspension passed the membrane. Based on this estimation, it is reasonable to conclude that complete blocking, which assumes that every particle seals a pore, could not be responsible for the flux decline throughout this entire period of the experiment. This conclusion leaves intermediate blocking as the mechanism possibly responsible for the initial flux decline raising the question of the availability of membrane surface pores for the particles to block. Given that the skin layer of the RO membrane is non-porous and the pores of NF membranes are, by the definition of nanofiltration, smaller than 5 nm (Fig. 3.4), the fact that intermediate blocking and not cake filtration described the experimental flux data the best is surprising and hints at the fortuitous nature of the fit. One possible physical explanation of the good fit by the intermediate blocking model is that some variance exists in the permeability of the membrane skin. In this case, more permeable areas (surface defects) [24] of the membrane skin would function as “pores” during initial stages of the separation experiment resulting in higher colloidal deposition on these areas. After the membrane surface is completely covered by the colloids, no “pores” on the membrane surface would remain to be blocked and only cake filtration would occur afterwards. This explanation is consistent with and corroborated by the reported short- 92 a BW30-365 . b NF90 (e) UF—300kDa Figure 3.4. SEM images of the filtration surfaces of membranes. (a) BW30-365; (b) NF90; (c) UF-30kDa; (d) UF-lOOkDa; and (e) UF-300kDa. 93 term increases in the rejection of salt by nanofiltration membranes immediately upon the introduction of colloids into the electrolyte-only feed [25, 26]. An additional, and not necessarily alternative, explanation of the deviation from the cake filtration law is the transient behavior of the cake porosity during the initial stages of the cake growth. As the estimation of the effective porosity using Carrnan-Kozeny equation demonstrates (Fig. 3.5), the effective porosity increased initially for a short period and then decreased until a relatively steady value was achieved; this implies that with the deposition of more particles an initially loose deposit grew into a more compact structure until a stable and uniform cake was formed. The surface roughness of BW30-365 and NF 90 membranes (Fig. 3.4) may be contributing to this transient behavior of cake porosity. While it is commonly assumed that for NF and R0 membranes cake filtration is the only mechanism of colloidal fouling, our results indicate that as much as 18% (for BW30-365) and 53% (for NF90) of flux decline for salt—rejecting membranes can not be unequivocally assigned to cake formation. Given that colloidal deposition contributing to the flux decline at the early stage of filtration is the likely cause of irreversible portion of the fouling, the reported findings point to the importance of more detailed understanding of the colloid-membrane interactions that manifest themselves and can be interpreted as pore blocking. 94 0.4 I I I I I I I I I l I I I I I I f I .0 (a) 1 1 g ”Elli ......... --'- [with . . e .0 ; fl . ~ é. 02 _ Q NF90 . .- ._ z - El BW30-365 ' E 4 , .1 a: I“ t - 0.1 r _ , i 1 1 " T 0 1 1 1 1 1 1 1 1 1 l 1 1 1 1 1 1 1 1 1 0 50 100 150 200 Permeate volume (mL) Figure 3.5. Effective porosity of the layer of silica colloids deposited onto the surface of BW30-365 and NF90 membranes. 95 3.4. Conclusions Experiments on the dead-end filtration of colloidal suspensions using membranes ranging from loose ultrafiltration to reverse osmosis were conducted to determine the type, the chronological sequence, and the relative importance of individual fouling mechanisms. To identify fouling mechanisms, a method was proposed that involves a straightforward linear least squares fitting with an easily determinable goodness of fit. It was demonstrated that 1) 2) 3) 4) For all membranes, flux decline during the later stages of filtration is consistent with the cake filtration mechanism, but deviates from this behavior during earlier stages of filtration. For all membranes, flux decline during the earlier stages of filtration could be interpreted in terms of one or more pore blocking mechanisms. The observed temporal overlap of different fouling mechanisms was interpreted as the transitional stage wherein contributions of the two fouling mechanisms to the flux decline were commensurate. For ultrafiltration membranes, pore blocking and not cake formation was the largest contributor to the observed flux decline. The chronological sequence of blocking mechanisms was interpreted to depend on the size distribution and surface density of membrane pores. For salt-rejecting membranes, flux decline during the earlier stages of filtration was attributed to either intermediate blocking of relatively more permeable areas of the 96 membrane skin, or to the cake filtration in its early transient stages, or a combination of these two mechanisms. Clear identification and differentiation between fouling mechanisms acting during the early stages of membrane separation provides useful insights into details of colloid- membrane interactions that are implicated in controlling permeate flux performance and the efficiency of membrane cleaning. Acknowledgements This work was supported by the National Science Foundation (research grant OISE- 0530174) and the National Water Research Institute (project no. 05-TM-007). We thank Dow-FilmTec for providing membrane samples and Nissan Chemical America Corp. for supplying the silica suspension. 97 Reference 10. ll. 12. 13. P. H. Hermans and H. L. Bredée, Principles of the mathematical treatment of constant-pressure filtration, J. Soc. Chem. Ind. 55T (1936) 1-4. V. E. Gonsalves, A critical investigation on the viscose filtration process, Rec. Trav. Chim. des Pays-Bas 69 (1950) 873. H. P. Grace, Structure and performance of filter media. II. Performance of filter media in liquid service, AIChE. J. 2 (1956) 316-336. J. Hermia, Constant pressure blocking filtration laws - application to power-law non-Newtonian fluids, Trans. Inst. Chem. Eng. 60 (1982) 183 -187. K. J. Hwang and T. T. Lin, Effect of morphology of polymeric membrane on the performance of cross-flow microfiltration, J. Membr. Sci. 199 (2002) 41-52. E. Iritani, N. Katagiri, Y. Sugiyarna and K. Yagishita, Analysis of flux decline behaviors in filtration of very dilute suspensions, AIChE J. 53 (2007) 2275-2283. H. Huang, T. A. Young and J. G. Jacangelo, Unified membrane fouling index for low pressure membrane filtration of natural waters: principles and methodology, Environ. Sci. Technol. 42 (2008) 714-720. W. Xu and S. Chellam, Initial stages of bacterial fouling during dead-end microfiltration, Environ. Sci. Technol. 39 (2005) 6470 -6476. E. Iritani, Y. Mukai, Y. Tanaka and T. Murase, Flux decline behavior in dead-end microfiltration of protein solutions, J. Membr. Sci. 103 (1995) 181-191. L. Palacio, C. C. Ho, P. Pradanos, A. Hernandez and A. L. Zydney, Fouling with protein mixtures in microfiltration: BSA-lysozyme and BSA-pepsin, J. Membr. Sci. 222 (2003) 41-51. C. C. Ho and A. L. Zydney, A combined pore blockage and cake filtration model for protein fouling during microfiltration, J. Colloid Interface Sci. 232 (2000) 389. W. Yuan, A. Kocic and A. L. Zydney, Analysis of humic acid fouling during microfiltration using a pore blockage-cake filtration model, J. Membr. Sci. 198 (2002) 51-62. J. A. Suarez and J. M. Veza, Dead-end microfiltration as advanced treatment for wastewater, Desalination 127 (2000) 47-5 8. 98 14. 15. l6. 17. 18. 19. 20. 21. 22. 23. 24. 25. M. K. Purkait, P. K. Bhattacharya and S. De, Membrane filtration of leather plant effluent: Flux decline mechanism, J. Membr. Sci. 258 (2005) 85—96. T. Mohammadi, M. Kazemimoghadarn and M. Saadabadi, Modeling of membrane fouling and flux decline in reverse osmosis during separation of oil in water emulsions, Desalination 157 (2003) 369-375. P. Aimar. Recent progress in understanding particle fouling of filtration membranes. in The Fifth International Membrane Science & Technology Conference. 2003. Sydney, Australia. W. R. Bowen, J. I. Calvo and A. Hernandez, Steps of membrane blocking in flux decline during protein microfiltration, J. Membr. Sci. 101 (1995) 153-165. C. Duclos-Orsello, W. Li and C. C. Ho, A three mechanism model to describe fouling of microfiltration membranes, J. Membr. Sci. 280 (2006) 856. E. M. Tracey and R. H. Davis, Protein fouling of track-etched polycarbonate microfiltration membranes, J. Colloid Interface Sci. 167 (1994) 104-116. K. J. Kim, V. Chen and A. G. Fane, Ultrafiltration of colloidal silver particles: Flux, rejection, and fouling, J. Colloid Interface Sci. 155 (1993) 347 - 359. S. F. E. Boerlage, M. D. Kennedy, M. R. Dickson, D. E. Y. El-Hodali and J. C. Schippers, The modified fouling index using ultrafiltration membranes (MFI-UF): characterisation, filtration mechanisms and proposed reference membrane, J. Membr. Sci. 197 (2002) 1-21. H. Choi, K. Zhang, D. D. Dionysiou, D. B. Oerther and G. A. Sorial, Effect of permeate flux and tangential flow on membrane fouling for wastewater treatment, Sep. Purif. Technol. 45 (2005) 68-78. D. C. Montgomery, A. E. Peck and G. G. Vining, Introduction to Linear Regression Analysis. Fourth ed. 2006: Wiley-lnterscience. B. Mi, C. L. Eaton, J H. Kim, C. K. Colvin, J. C. Lozier and B. J. Marinas, Removal of biological and non-biological viral surrogates by spiral-wound reverse osmosis membrane elements with intact and compromised integrity, Water Res. 38 (2004) 3821-3832. B. M. Hoek and M. Elimelech, Cake-enhanced concentration polarization: a new fouling mechanism for salt-rejecting membranes, Environ. Sci. Technol. 17 (2003) 5581-5588. 99 26. F. Wang and V.V. Tarabara, Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance, J. Membr. Sci. 293 (2007) 111-123. 100 SS. SUPPLEMENTARY DATA 025 .fi.,...,-..,..-,...,.-.‘ 0,26...-,-...,--..,s...,...MWA 1 ' i : 5 . L —' IO.138+0.00436X .R- 0.998 1 ; —y=o.~1o4+o.o111x; as 0.997 1‘): . 1' ; s s , 1b). 0.225 2. ., , . .1 _. 024 _._ . 1 "gt a - ~ . 1 ‘ t - . g ' --" : ‘ 9 0.2 _- 30.22 - ~7 E : i; l 1 20.175 - °- 02 - - .. E . e .. . Q 0.15 — 3 0.18 - i 1- ‘ < 0.125 1 0.16 — - 1 1 * 0.1[1L_1_l111i1_1111_t1_1111111114 0.14”111114441111L1141111111111111 0 2 4 6 8 10 12 0 5 10 15 20 25 30 Time (min) Time (min) (13 ."* . , . . . , . - . , . . a , . . .+ - 3 1 (C) : —y80.0s746+0.000926x th1 1 0.27 s ; .. .. 1_ .. 1. ..... ,2 :1‘ E 0.24 7 E 1 5. : §;(121 - 0.18 :- 0'15 ’ 1 1 1 1 1 1 1 1 1 . . 1 1 1 1 1 1 1 ‘ 100 120 140 160 180 200 Permeate volume (mL) Fig. S3.]. Application of blocking laws to the filtration data of UFlOOkDa membrane: (a) 1/ Q versus t based on Eq. (3.7) for intermediate blocking; (b) t/ V versus t based on Eq. (3.2) for standard blocking; and (c) t/V versus V based on Eq. (3.4) for cake filtration. 101 0.6 . y= 0.473 + 0.000402x R= 0.999 1 . ""'""Y=0.423+0.00(i569x R=1§ 1 0.55 9 = = LT g . .5 0.5 g 1 g 1- 0.45 1 i +ewao-365 ’ "0—NF90 0.4 b 1 1 1 1 l 1 1 1 1 l 1 1 1 l 1 1 1 a 0 50 100 150 200 Permeate volume (mL) Fig. 83.2. Application of blocking laws to the R0 and NF membrane filtration data: t/V versus V based on Eq. (3.4) for cake filtration. 102 I fii I r I ' l 2,3 4 1 1 , . 1 1 1 y = 1.98 - 0.001865)! R= 0.9901 : —El— BW30-365 Z ..1 Q (cm‘3lmin) N .x (D T I I I ..s m 80 0 20 40 60 100 Permeate volume (mL) 0.6.........,..-2,fi,.,,..,, f (b)-—-y=0.5053+0.0009971x R= 0.9901 i ‘ y=o.4404+0.00219x Rs 0.9987 0.55 — 1 . ; _ _ 6‘ < 5 . a 0 5 ~ E 1 E I NF90 +BW30-365 =0 45 1 —e—Ni=9o . ‘ i 1 T 04 11111111111111111111 0 10 20 30 40 50 Time (min) Fig. 83.3. Application of blocking laws to the R0 and NF membrane filtration data: (a) Q versus V based on Eq. (3.6) for complete blocking, and (b) 1/ Q versus t based on Eq. (3.7) for intermediate blocking. 103 As Fig. S3 shows, both Eq. (3.6) and Eq. (3.7) fit the initial part of flux data for both R0 and NF membranes well. The underlying reason is explained as follows. K b in Eq. (1) and K ,- in Eq. (3.3) were calculated numerically based on the V(t) data obtained during initial stages of R0 and NF experiments. It is found that for both the R0 and NF membranes, the value of K b is on the order of 10'16 or 10'17 min", and the value of K ,- is on the order of 10'17 cm'3. This means that Kbt << 1 and K iQOt << 1. —K b1 For Eq. (1) which corresponds to complete blocking, K b V = Q0 (1 — e ), Taylor series K312 K {:73 . . . . . 2 + 6 ).D1fferent1at1ngth1s equatlon leads expansion gives us, K bV z Q0 (K bt — IO: 2 2 dV Kbt =—~ l—K 1+ . (3.1) Q dt Q0( b 2 ) For Eq. (3) which corresponds to intermediate blocking, K ,-V = ln(1+ K iQOt) , a similar approach gives us, Q = % ~ Q0(1—K1Q61+K12Q3t2)- ($32) The similarity between Eq. (83.1) and Eq. ($3.2) demonstrates that, when K bt << 1 and K iQot << 1, Eqs. (3.1) and (3.3) predict the same relationship between flux and time. This is the reason why both Eqs. (3.6) and (3.7) fit the initial part of flux data equally 104 well. This implies that, when K bt << 1 and K iQot << 1, Eqs. (3.6) and (3.7) can not be used to discriminate between complete blocking and intermediate blocking. 105 CHAPTER FOUR Wang, Fulin; Tarabara, Volodymyr V. (2008). Modeling the permeability of fiber-filled porous media using Kozeny-Carman-Ethier equation. (In preparation for publication in a journal) 106 CHAPTER FOUR MODELING THE PERMEABILITY OF FIBER-FILLED POROUS MEDIA USING KOZENY-CARMAN-ETHIER EQUATION ABSTRACT In this study, a model is proposed for predicting the permeability of mixed porous media of particles and fibers. The model combines Kozeny-Carman equation and Ethier’s theoretical model to predict the permeability of the mixed media based on the permeability of the porous bed of particles and on the volume fraction of the fibers. Membrane filtration experiments were conducted using silica colloids and carbon fibers with varied fiber mass fraction. The model prediction is in great accordance with the experimental results for the fiber mass fraction of less than 50%. 107 4.1. INTRODUCTION Flow through porous media is a very important process for a variety of natural phenomena and engineering applications, such as, flow of groundwater through soil, water transport through bodily tissues, flow through deposit layers on membrane surfaces, flow through fuel cells and permeation through polymeric networks. Given the kaleidoscopic shapes of natural and engineering matter, the components of the media can be spherical particles, elongated particles, rods, cubes, fibers (cylinders), or a combination of them. Permeability is the key parameter for designing and controlling this process. The permeability depends on the structure of the media, such as, the shape of the components, the porosity, the packing densities, etc [1, 2]. Based on Darcy’s law, theoretical equations have been developed for predicting the permeability of porous media. For media of uniform spherical particles, the well-known Kozeny-Carman equation [3] is widely used, which predicts the permeability (K) based on porosity (e) and particle diameter (dp ) as shown in Eq. (4.1), d283 ___ P 2 . (4.1) 180(1 — e) Darcy’s law has also been extended to model the permeability of fibrous media. The flow through media of fibers with uniform radius that are placed randomly or in arrays has been modeled by several researchers [4-8]. Jackson and James [2] summarized the historical development of this subject. A general equation can be presented as [1, 2], 108 K —2 = f (it). (4.2) a where a is the fiber radius, f ((25) is a model-dependent function, and ¢ is the fiber volume fraction. However, it is only applicable for monodisperse fibrous media, but not polydisperse fibrous media. Ethier [1] theoretically modeled the flow in a composite porous media composed of two types of fibers of highly dissimilar sizes using the Debye-Brinkman equation. The overall permeability was determined based on the permeability of the fine fibers and the volume fraction of the coarse fibers (more detail in Section 4.2.3). Fibrous filters have been widely used in industries such as production of chemicals, environmental safety, food industry, pharmaceutical, and semi-conductor industries ([9, 10]) to remove gas impurities. When solid particles accumulate inside the filters, the filters turn into a matrix of particles and fibers. Thomas et al. [10] modeled the performance of fibrous filters of varied characteristics removing solid aerosol particles in air under different operating conditions. Pressure drop and filter efficiency were modeled based on cake packing density. In this study, the trapped particles formed dendrites and functioned as new collecting fibers. Recently, new nanofibrous membranes have been developed using various materials, such as TiOz nanowire [ll], polyurethane [12], polyvinyl alcohol [13, 14], polysulfone [15], and nylon [16]. As demonstrated in [15, 16], when the nanofibrous membranes are fouled by particles, they become filtration media of 109 fibers and particles. In media of fibers and particles, fibers increase tortuosity and obstruct some permeable area, which greatly reduces the total permeability [1]. Knowledge of the permeability of such media is crucial for designing membrane systems and predicting their performance. However, to our best knowledge, so far there have been no theoretical studies on predicting the hydraulic permeability of porous media consisting of particles and fibers. To predict the performance of these media is in pressing need. In this study, we developed a theoretical model for predicting the hydraulic permeability of media consisting of particles and fibers, based on Kozeny-Carman equation and Ethier’s theoretical model. The model prediction is in great accordance with the experimental results. 4.2. Theoretical background 4.2.1. The hydraulic resistance of a deposit layer in membrane systems When the feed water is free of solutes and potential foulants, the pure water permeate flux J 0 is linked to the transmembrane pressure AP by the Darcy’s equation: JO :——, (4.3) 110 where Rm is the hydraulic resistance of the membrane, and ,u is the dynamic viscosity of water. When there are foulants deposited on membrane surface, the permeate flux will decrease due to the hydraulic resistance of the deposit layer (Rd ). The equation for permeate flow becomes: : AP #(Rm '1' Rd (1)) , 1 (4.4) where J, is the permeate flux at filtration time t, and Rd (t) is the hydraulic resistance of the deposit layer as a function of t. J , /J0 is defined as the normalized permeate flux. When filtration experiments are conducted at a constant transmembrane pressure, Eqs. (4.3) and (4.4) lead to a simple equation for determining Rd (t) as: Rd (1) = [JOJ’ J’ :lRm. (4.5) 1 Equation (4.5) can also be used for determining the Rd as a function of permeate volume (Rd (V) ) as: Rd (V) = [JOJ‘VJV JR," , (4.6) 111 where J V is the permeate flux when the cumulative permeate volume is V . 4.2.2. The hydraulic resistance of porous media of particles For deposit layers consisting of spherical incompressible particles of uniform size, Kozeny-Carman equation [3] is widely used to calculate Rd : 180(1—s) Mp ,apdfir;3 A Rd = 1 (47) where A is the membrane filtration area, a is the porosity, M p is the mass of particles deposited on membrane surface, and tip and p p are the diameter and the density of the particles, respectively. The specific hydraulic resistance lid is defined as: .1 R _ 2 R __d_=180(1£) — 2 4.8 5d die3 ( ) where 6d is the thickness of the deposit layer. The permeability K is the reciprocal of the hydraulic specific resistance, as K = 1/ Rd. 4.2.3. The permeability of highly polydisperse fibrous media Ethier [1] theoretically modeled the flow in a composite porous media composed of two types of fibers of highly different sizes using the Debye-Brinkman equation. The fine 112 matrix appeared to be homogeneous on the length scale of the coarse fibers. In his model, the coarse fibers were embedded into the fine fibers (with permeability K f) to form a porous matrix, as shown in Figure 4.1. Ethier has shown that, for a medium of randomly oriented fibers, when K f / ac2 -—> 0 where ac is the radius of the coarse fibers, the total permeability (K m) can be predicted based on the permeability of the fine fibers (K f) and the volume fraction of the coarse fibers as: Ktotz 1—¢c (49) Kf 1+2/3¢c’ ' where 05, is the volume fraction of the fraction of the coarse fibers, which is defined as: _ VC Vtot , ¢c (4.10) where VC is the volume occupied by the coarse fibers, and Vm, is the total volume. 4.3. Experimental method 4.3.]. Experimental apparatus Dead-end membrane filtrations were conducted using a high-pressure stainless steel filtration cell (HP 4750, Sterlitech Corp., Kent, WA) at a transmembrane pressure of 100 113 psi (0.689 MPa) without stirring. The permeate flux was recorded using a digital weighing balance (ARC120, Ohaus Corp., Pine Brook, NJ) interfaced with a computer via a built-in RS-232 port. The permeate was delivered to a beaker positioned on the balance. The data from the balance were logged to the computer via a program written in LabView (version 7.1, National Instruments, National Instruments Corp., Austin, TX) with a time interval of 10 s. The volume of the feed suspensions was 100 mL. The experiment was conducted at the room temperature (21 °C). Because there was no stirring in the filtration cell, a constant temperature of the feed suspension was assumed. 114 Figure 4.1. [1] Idealized cross-section view of the porous matrix (The coarse fibers are shown to be parallel to one another in this figure, but they can be randomly oriented as well) 115 4.3.2. Reagents and chemicals All reagents were of ACS analytical grade or higher purity (Fisher Scientific, Pittsburgh, PA) and were used without further purification. The deionized water used in the experiments was supplied by a commercial ultrapure water system (Lab Five, USFilter Corp., Hazel Park, MI) equipped with a terminal 0.2 pm capsule microfilter (PolyCap, Whatman Plc., Sanford, ME). The resistivity of the water was greater than 16MQ-cm. The pH of water was measured to be in the range of 6.1 to 6.4 (pH meter D12, Horiba, Kyoto, Japan). No buffers were added to the water. 4.3.3. Ultrafiltration (U F) membrane, silica colloids, and carbon fibers A polyethersulfone (PES) UF membrane that has a nominal molecular weight cut-off (MWCO) of 10 kDa according to the manufacturer was used (DiaryUF, Hydranautics Corp., Oceanside, CA). This UF membrane was chosen because it did not reject salts and thus no possible osmotic pressure was expected to occur across the membrane. The membrane was supplied as a flat sheet and was stored in a sealed plastic container at room temperature. Before the experiment, two rectangular membrane coupons were cut out of the membrane sheet and soaked in deionized water for 24 h at 4 °C with water exchanged after the first 12 h. Silica colloids (SnowTex-XL, Nissan Chemical America Corp., Houston, TX) were used as the model colloids. Carbon fibers were provided by Applied Sciences, Inc. (Pyrograf® PR-19-PS-LD, Cedarville, Ohio) with “PS-LD” indicating that the fibers were low density (LD) pyrolytically stripped (PS) carbon fibers. 116 4.3.4. Scanning electron microscopy (SEM) analysis The colloids, the carbon fibers, and the deposit layers on membrane surface were first coated for 20 s using a pure osmium coater (Neoc-AN, Meiwa Shoji Co. Ltd., Kyoto, Japan), and then imaged using Hitachi S-47OOII Field Emission scanning electron microscope (Hitachi High Technologies America, Inc. Pleasanton, CA) with a secondary electron detector. The microscope was operated in ultra-high resolution mode with an accelerating voltage of 15 kV and an emission current of 10 11A. 4.3.5. Experimental protocol 4.3.5.1. Preparation of the feed suspensions Carbon fibers of were weighed at 200 mg and added to a 100 mL 2 mM sodium dodecyl sulfate (SDS) solution. The suspension was continuously sonicated with a 100 W ultrasonic cell disrupter (VirSonic 100, SP Industries Inc., Gardiner, NY) for 40 min. After that the fibers were evenly and stably dispersed in the suspension. After 24 hours, no visible aggregates were observed. The suspension was used as the stock suspension of the carbon fibers. 4.3.5.2. Membrane compaction To eliminate the effect of membrane compaction on our determination of hydraulic resistance of deposit layers based on permeate flux, two UF membrane coupons were compacted at a transmembrane pressure of 120 psi (0.827 MPa) using deionized water in a crossflow filtration system for around 48 h until a stable permeate flux was achieved. 117 The crossflow filtration system was described in our previous study [17]. After the compaction, the membrane coupons were taken out and several round pieces (with a diameter of 47.6 mm) were gently cut out of the coupons. The pieces were suitable for the dead-end filtration cell and were stocked in deionized water at 4 °C for subsequent use. 4.3.5.3. Measurement of membrane hydraulic resistance using pure water The permeate flux was recorded for a series of transmembrane pressures: 60 psi (0.414 MPa), 80 psi (0.552 MPa), 100 psi (0.689 MPa) and 120 psi (0.827 MPa). The hydraulic resistance of the clean membrane (Rm) was calculated from Eq. (4.1) using linear least squares fitting. 4.3.5.4. Membrane filtration experiments One hundred milliliters of feed suspensions of varied concentration of carbon fibers and 200 mg/L silica colloids were prepared using the stock suspensions of fibers and silica colloids. The feed suspensions were sonicated for 5 minutes using a sonicator (Model SOT, VWR International Inc., West Chester, PA) to ensure the fine dispersion of silica colloids and carbon fibers. After that the feed suspensions were added into the filtration cell and membrane filtration experiments were conducted at a constant transmembrane pressure of 100 psi (0.689 MPa). The permeate flux was recorded. The resistance of the deposit layer was calculated based on Eqs. (4.5) and (4.6) as a function of time and permeate volume as well. 118 4.4. Results and discussion 4.4.1. Characterization of the silica colloids and carbon fibers The SEM image of the silica colloids is shown in Fig. 4.2(a). The colloids have a density of 2.36 g/cm3 and a diameter of (55 .6 i 1.3) nm, as characterized in our previous study [17]. Fig. 4.2(b) shows the SEM image of the carbon fibers. According to the manufacturer, the fibers have a density of 1.95 g/cm3 and the diameter and the length range from 0.1 to 0.2 pm and from 30 to 100 pm, respectively. 119 5 .3) \‘ r 1 'V (, Y3" ‘3 (I ..., r J k 1 k ’1 %“P~f3oofl~1v:s 050 r-C 36 1‘0: (a) (b) Figure 4.2. SEM images for the silica colloids (a) and carbon fibers (b) 120 fifififllifli‘ifl . r 9 co .0 a: 3’40. - '- 4.0.: mums-we, . '::.1 fi-‘(o , ' ' x1 , "nu/.1 “r ’0’,” i '92.}? (((‘p‘b/l “ Normalized permeate flux .0 .O O) \l 9 100 mgIL carbon fibers 0 50 mglL carbon fibers 0 5 A 25 mg/L carbon fibers _______ .............. _: O 200 mglL carbon fibers 3 . D Silica particles only g 1 0 4 £21 200 mgIL carbon fiber only i 0 20 40 60 80 100 Permeate volume (mL) 1 , . . . r . 0 100 mgIL carbon fibers § § .% O 50 mglL carbon fibers 5 g.-.-:-:. ... A 25 mgIL carbon fibers 3 ._ . 5 0-8 0 200 mglL carbon fibers """"""" """""" - . k M D Silica particles only 5 . -- 3 P2 fl. . 8 05 .............. c 8 .2 In 8 0.4 .2 3 E 'a .. 3. I 02 ................................................. O O 20 40 60 80 100 Permeate volume (mL) Figure 4.3. Permeate flux (a) and hydraulic resistance of the deposit layer (b). Transmembrane pressure was 100 psi (0.689 MPa) for each experiment. Rm in each experiment was shown in Table 1. All of the experiments were conducted with 200 mg/L silica colloids, except the data line for “200 mg/L carbon fibers only” in (a) when no colloids were added. The hydraulic resistance of the deposit layer from “200 mg/L carbon fibers only” was not shown in (b) because it was too low to be measurable in our experiments. 121 4.4.2. Permeate flux and hydraulic resistance of the deposit layer Table 4.1 shows the membrane hydraulic resistance (Rm) in the experiments. Rm was used to calculate the hydraulic resistance of the deposit layer (Rd) based on Eqs. (4.5) and (4.6). The normalized permeate flux of the UP membrane filtrating different suspensions is shown in Fig. 4.3(a). All of the experiments were conducted with 200 mg/L silica colloids, except the experiment with carbon fibers only when no silica colloids were used. From Fig. 4.3(a) it can be seen that no evident permeate flux was observed in the experiment with 200 mg/L nanofibers only, which indicates the fibers formed a very porous structure that had a relatively low hydraulic resistance to the membrane hydraulic resistance. SEM images (Fig. 4.4(e)) confirmed that the deposit layer from this experiment was quite porous with visible large open spaces among the fibers. With 200 mg/L silica colloids, the permeate flux decline rate increased when the concentration of the carbon fibers increased from 25 mg/L to 100 mg/L (Fig. 4.3(a)), indicating the hydraulic resistance increased with increased carbon fiber concentration (Fig. 4.3(b)). However, when it reached 200 mg/L, permeate flux decline rate greatly decreased, indicating a lower hydraulic resistance. This is due to the fact that in this experiment the deposition layer was more porous with open spaces among the fibers, as shown in Fig. 4.4(d), which dramatically reduced the hydraulic resistance. 122 Table 4.1. Rm in different experiments Experiment Name R m (x1olym'1) 200 mg/L carbon fibers only 1.17 200 mg/L silica colloids only 1.16 25 mg/L carbon fibers + 200 mg/L silica colloids 1.22 50 mg/L carbon fibers + 200 mg/L silica colloids 1.22 100 mg/L carbon fibers + 200 mg/L silica colloids 1.16 200 mg/L carbon fibers + 200 mg/L silica colloids 1.43 123 It is worthy noticing that in Fig. 4.3(a), the permeate flux from the experiment with 200 mg/L silica colloids only decreased slightly faster than that with 200 mg/L silica colloids and 200 mg/L carbon fibers. But the hydraulic resistance of the deposit layer for the former was slightly lower than that for the latter (Fig. 4.3(b)). This is due to the lower hydraulic resistance of the membrane used in the former, as shown in Table 4.1. 4.4.3. SEM images of the deposit layers 4.4.4. Modeling the hydraulic resistance of the deposit layers consisting of silica colloids and fibers 4.4.4.1. Theoretical development In our experiments, the fibers (with a radius a f = 0.05 ~ 0.1 pm) and silica colloids can be regarded as the coarse and fine fibers in Ethier’s theoretical model, respectively. Because in our experiments K p M} < 0.001 where K p is the permeability of the porous bed formed by the silica colloids, Ethier’s model can be used to predict the permeability of the deposit layer based on the permeability of the silica colloid aggregates. Replacing the permeability K by hydraulic specific resistance Ii , Eq. (4.9) can be re-written as: Riot =1+2/3¢f (411) RP 1—¢f ' 124 Figure 4.4. SEM of the deposition layers (continued on next page). 125 (0) Figure 4.4. SEM of the deposition layers. (a) 25 mg/L carbon fibers and 200 mg/L silica colloids; (b) 50 mg/L carbon fibers and 200 mg/L silica colloids; (c) 100 mg/L carbon fibers and 200 mg/L silica colloids; (d) 200 mg/L carbon fibers and 200 mg/L silica colloids; (e) 200 mg/L carbon fibers only (without silica colloids) 126 where 13,0, is the hydraulic specific resistance of the total deposit layer, 11,, is the specific hydraulic resistance of porous bed of the silica colloids, and eff is the volume fraction of the carbon fibers. Combining Eqs. (4.7), (4.8) and (4.11), an equation for predicting the hydraulic resistance of a deposit layer consisting of fibers and particles can be written as: . lsoaf-s )2 r+213¢ Rd = Rdé‘d = 2 5 [ f ]§d, (4.12) dpep 1—¢f where 6d is the thickness of the deposit layer and 8,, is the effective porosity of the porous bed of silica colloids. We name this equation as Kozeny-Carman—Ethier equation. The total volume of the deposit layer V can be calculated as: (0! mp mf + ’ pp(1_8p) pf V10: = Vp + Vf = (4.13) where Vc and Vp are the volume of carbon fibers and the porous bed of silica colloids on membrane surface, respectively, p f and p p are the density of the carbon fibers (1.95 g/cm3) and the silica colloids (2.36 g/cm3), respectively. The thickness of the deposit layer (5d ) can be calculated as: 127 mp ’"f + 5d = V10. = ”PO-Sp) pf : ”'pr +mfpp0‘8p) (4.14) A A Apfppfl-ep) ’ where A is the area of the deposit layer (i.e. the effective filtration area of the membrane). The volume fraction of the carbon fibers ¢f is: m _L V p m p (1-6 ) ¢f=Vf = f = f P P . (4.15) to, mf+ mp mfpp(1-£p)+mppf pf pp(l-6p) Combining Eqs. (4.12), (4.14) and (4.15), the equation for predicting the hydraulic resistance of the deposit layer consisting of carbon fibers and silica colloids in our experiments is: (4.16) R 180(1—6p) 3mppf+5mfpp(1-£p):|[mppf+mfpp(l-£p) d= - dg£p3 3’"pr Apfpp 4.4.4.2. The effective porosity of the porous bed of silica colloids We assumed that the back transport of silica colloids from membrane surface to the feed was negligible, which means that the mass of the silica colloids deposited can be calculated based on permeate volume and the feed concentration of the silica colloids. This assumption is reasonable because the filtration was performed using a dead-end 128 mode without stirring. This assumption has been widely used in the models for the performance of membrane operated at a dead-end non-stirring mode. For examples, it is used in the well-know blocking filtration laws [18] which study different blocking mechanisms in membrane fouling processes. Thus, the mass of silica colloids deposited on the membrane surface can be predicted based on the permeate volume (V ) and the feed concentration of the colloids (C p ), as: p. (4.17) Based on Eqs. (4.6), (4.7) and (4.17), the efi‘ective porosity of the deposit layer of silica colloids only can be determined as a function of permeate volume, as shown in Fig. 4.5. It can be seen that the effective porosity gradually increased until it reached a stable value, which is believed to be attributed to the processes of breakup and restructuring of the deposit layer. The averaged effective porosity was 0.353 based on over 130 data points. This average porosity was used for the following modeling analysis. 129 0.8 V l I I r r I I l v I I r r r l 0.6 E 3 8 $0.4 § 3: Ill 0.2_ ....................................................................................... __ J + _ , . ‘ 4 041.111L11111...i 0 20 40 60 80 100 Permeate volume (mL) Figure 4.5. Porosity of the silica deposit layer as a function of permeate volume 130 4.4.4.3. Prediction of the hydraulic resistance of the deposit layer consisting of silica colloids and carbon fibers We made four assumptions for applying the Kozeny-Carman-Ethier equation on our experimental data: (1) The back transport of silica colloids from membrane surface to the feed was negligible, as discussed above. This assumption enabled us to calculate the mass of carbon fiber and silica colloids on membrane surface based on permeate volume, as indicted in Eq. (4.17). (2) The porosity of the silica aggregates did not change in the presence of carbon fibers. This is a valid assumption because all of the experiments were operated at dead-end non-stirring mode with the same solution chemistry. As suggested in the literature [17, 19-21], solution chemistry (i.e. ionic strength and pH) is the main factor controlling the structure of deposit layers on membrane surfaces because it affects the surface charge of the colloids and thus influences the interactions among the colloids. As mentioned above, the averaged porosity (0.353) was used for the bed of silica colloids in the mixed media. (3) The carbon fibers and colloids are uniformly dispersed in the deposit layers. This is ensured in our experimental approaches. As shown in Section 4.3.5.1, the carbon fibers were stably and evenly dispersed after the intense sonication. In addition, immediately before each filtration experiment the feed suspension was sonicated again for 5 minutes to achieve a well-mixed suspension. 131 (4) In the mixed media, all of the open spaces between the fibers are occupied by the colloids. This is the assumption made in Ethier’s approach [1], which ensures that the permeate flow occurs among particles. As the SEM images in Fig. 4.4 suggests, this assumption was valid for the deposit layers formed with 25 to 100 mg/L fibers with 200mg/L silica colloids (Fig. 4.4(a-c)), but was not valid for the 200 mg/L carbon fibers with the silica colloids Fig. 4.4(d) where there were open spaces among the fibers. Based on these assumptions, Rd , which was theoretically predicted using Eq. (4.16), was compared with experimental measurement from Eq. (4.6). The results are shown in Fig. 4.6. From Fig. 4.6, we can see that the model prediction was very good for the deposit layers from 25 mg/L, 50 mg/L and 100 mg/L carbon fibers with 200 mg/L silica colloids. However, it became not applicable when the carbon fiber concentration was increased to 200 mg/L. This is because there were open spaces among the fibers (Fig. 4.4(d)), which violated the fourth of the aforementioned assumptions. The open spaces greatly decreased the hydraulic resistance of the deposit layer. 132 1 . .1 ..s,. 14 .4., 4 4. ............. P a E] EXPO” “I Musunmm ..... ............... .1 0.8 E] 0 Model prediction «114-7 Experimental measurement 0 Model prediction .0 a: T I 0.6 : LLJ ' t 0.4 .......... ...... . ...... 0.4 0.2 0.2 ' .2 22’ t z j , (c) . ' (d) : 1 Cl Experimental measurement “ I j 2 i ‘ ‘ ------------ .. ‘P 0 Model prediction ‘ D Experimental measurement . : 3)“; 1 0 Model prediction 0.3 _ ........................................................ in, ............... a 1.6 3 : Deposition layer resistance (10“ m") Deposition layer resistance no” 111") . O 0.6 ~ 1.2 eeeee 0.4 J 0.8» , . . j ‘ ”1,! -' .' .7 '. 1 : 0 I... . r . r . . A r A 1 r n A 0 . . . r . A . . r r L 0 20 40 60 80 100 0 20 40 60 80 100 Permeate volume (mL) Permeate volume (mL) Figure 4.6. Comparison between model prediction and experimental measurement for the hydraulic resistance of the deposition layer from the experiments with: (a) 25 mg/L carbon fibers and 200 mg/L silica colloids ( p2 = 0.92); (b) 50 mg/L carbon fibers and 200 mg/L silica colloids ( ,02 = 0.98); (c) 100 mg/L carbon fibers and 200 mg/L silica colloids (,02 = 0.97); (d) 200 mg/L carbon fibers and 200 mg/L silica colloids. 133 4.5. Conclusions In this study, we presented a Kozeny-Carrnan—Ethier equation to predict the permeability of a deposit layer consisting of fibers and particles. The equation was based on the permeability of the porous bed of silica colloids and the volume fraction of fibers. Dead- end membrane filtration experiments were conducted to experimentally determine the permeability of porous media with varied mass fraction of carbon fibers and silica colloids. The model prediction is in great accordance with the experimental results when the mass fraction of carbon fibers was less than 50%. When the mass fraction of carbon fibers exceeded 50%, the model was not applicable because the fibers created some open spaces, which were not occupied by particles, and thus greatly increased the permeability of the deposit layer. 134 Acknowledgments Financial support of this work by the National Water Research Institute (project no. 05- TM-007) is gratefully acknowledged. We thank Nissan Chemical America Corp. for supplying the silica suspension, Hydranautics Corp. for the ultrafiltraion (UF) membrane, and Applied Sciences, Inc. for the carbon fibers. We also thank Ms. Ewa Danielewicz from the Center for Advanced Microscopy and Dr. Jue Lu from the Composite Materials and Structures Center at Michigan State University for their assistance with osmium coating of samples for SEM imaging and preparation of carbon fiber suspensions, respectively. 135 Nomenclature fiber radius (m'l) radius of coarse fibers (m'l) radius of fibers in deposit layers consisting of particles and fibers (m'l) membrane filtration area (m2) feed concentration of silica colloids (kg/m3) diameter of particles (m']) a model-dependent function volumetric permeate flux (ms'l) volumetric pure water permeate flux (ms'l) volumetric pure water permeate flux at filtration time t (ms'l) volumetric pure water permeate flux at cumulative volume V (ms") permeability (m2) permeability of fine fibers (m2) permeability of porous bed formed by the silica colloids (m2) permeability of total media (m2) mass of particles in deposit layers consisting of only particles (kg) mass of fibers in deposit layers consisting of particles and fibers (kg) mass of particles in deposit layers consisting of particles and fibers (kg) hydraulic resistance of membranes to pure water (m'l) hydraulic resistance of deposit layers (m’l) specific hydraulic resistance (m'z) filtration time (h) cumulative permeate volume (m3) volume occupied by coarse fibers (m3) volume occupied by fibers in deposit layers consisting of particles and fibers (m3) volume occupied by particles total volume of deposit layers consisting of particles and fibers (m3) transmembrane pressure (MPa) 136 Greek letters 1” dynamic viscosity of water (kgm'ls'l) 5d thickness of the deposit layer (m) 6 effective porosity of deposit layer 5 p effective porosity of porous bed formed by silica colloids (15 volume fraction ¢c volume fraction of coarse fibers ¢f volume fraction of fibers in deposit layers consisting of particles and fibers pp particle density (kgm'3) p f fiber density (kgm'3) p2 correlation coefficient 137 Reference 10. ll. 12. 13. C. R. Ethier, Flow through mixed fibrous porous materials, AIChE J. 37 (1991) 1227-1236. G.W. Jackson and DP. james, The permeability of fibrous porous media, Can. J. Chem. Eng. 64 (1986) 364-374. P. C. Carman, Fluid flow through a granular bed, Trans. Inst. Chem. Eng. London 15 (1937) 150-166. J. Happel, Viscous flow relative to arrays of cylinders, AIChE J. 5 (1959) 174- 177. L. Spielman and Goren S.L., Model for predicting pressure drop and filtration efficiency in fibrous media, Environ. Sci. Technol. 279-286 (1968). GP Yu and T.T. Soong, A random cell model for pressure drop prediction in fibrous filters, Trans. ASME J. Appl. Mech. 301 (1975). S. Kuwabara, The forces experienced by randomly distributed parallel circular cylinder or spheres in viscous flow at small Reynolds numbers, J. Phys. Soc. of Japan 14 (1959) 527-532. A. S. Sangani and A. Acrivos, Slow flow past periodic arrays of cylinders with application to heat transfer, Int. J. Multiphase Flow 1982 (1982) 193-206. P.C. Raynor and D. Leith, Evaporation of accumulated multicomponent liquids from fibrous filters, Ann. Occup. Hyg. 43 (1999) 181 - 192. D. Thomas, P. Penicota, P. Contala, D. Leclerca and J. Vendelb, Clogging of fibrous filters by solid aerosol particles Experimental and modelling study, Chem. Eng. Sci. 56 (2001) 3549 - 3561. X. Zhang, A. J. Du, P. Lee, D. D. Sun and J. O. Leckie, Ti02 nanowire membrane for concurrent filtration and photocatalytic oxidation of humic acid in water, J. Membr. Sci. 313 (2008) 44-51. C. Yao, X. Li, K. G. Neoh, Z. Shi and E. T. Kang, Surface modification and antibacterial activity of electrospun polyurethane fibrous membranes with quaternary ammonium moieties, J. Membr. Sci. 320 (2008) 259-267. Y. Hong, T. Shang, Y. Li, L. Wang, C. Wang, X. Chen and X. Jing, Synthesis using electrospinning and stabilization of single layer macroporous films and fibrous networks of poly(vinyl alcohol), J. Membr. Sci. 276 (2006) 1-7. 138 14. 15. 16. 17. 18. 19. 20. 21. Y. Wang and Y. L. Hsieh, Immobilization of lipase enzyme in polyvinyl alcohol (PVA) nanofibrous membranes, J. Membr. Sci. 309 (2008) 73-81. R. Gopal, S. Kaur, C. Y. Feng, C. Chan, S. Ramakrishna, S. Tabe and T. Matsuura, Electrospun nanofibrous polysulfone membranes as pre-filters: Particulate removal, J. Membr. Sci. 289 (2007) 210-219. D. Aussawasathien, C. Teerawattananon and A. Vongachariya, Separation of micron to sub-micron particles from water: Electrospun nylon-6 nanofibrous membranes as pre-filters, J. Membr. Sci. 315 (2008) 11-19. F. Wang and V.V. Tarabara, Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance, J. Membr. Sci. 293 (2007) 111-123. J. Hermia, Constant pressure blocking filtration laws - application to power-law non-Newtonian fluids, Trans. IChemE. 60 (1982) 183 -187. W. R. Bowen, A. Mongruel and P. Williams, Prediction of the rate of cross-flow membrane ultrafiltration: A colloidal interaction approach, Chem. Eng. Sci. 51 (1996) 4321-4333. R. S. F aibish, M. Elimelech and Y. Cohen, Effect of interparticle electrostatic double layer interactions on permeate flux decline in crossflow membrane filtration of colloidal suspensions: An experimental investigation, J. Colloid Interface Sci. 204 (1998) 77-86. V. V. Tarabara, I. Koyuncu and M. R. Wiesner, Effect of hydrodynamics and solution chemistry on permeate flux in crossflow filtration: Direct experimental observation of filter cake cross sections, J. Membr. Sci. 241 (2004) 65-78. 139 CHAPTER FIVE Wang, Fulin; Tarabara, Volodymyr V. (2008). Crystallization of gypsum in the presence of colloidal silica. I: in a stirred batch system. (In preparation for publication in a journal) 140 CHAPTER FIVE CRYSTALLIZATION OF GYPSUM IN THE PRESENCE OF COLLOIDAL SILICA. I: IN A STIRRED BATCH SYSTEM ABSTRACT Although both scaling and colloidal fouling in membrane filtration processes have been intensively studied independently in their respective fields, there is very limited research in the context that these two fouling mechanisms occur simultaneously. In practice, fouling often involves different mechanisms due to the fact that feed waters usually have complex components. In this work, we studied the effect of silica colloids on gypsum scaling in a batch system using supersaturated calcium sulfate suspensions of varied saturation index in the absence and presence of silica colloids. The induction time was determined based on the light transmittance at 400 nm. Measuring conductivity of the feed suspensions was used as a convenient technique to determine the crystallization rate of gypsum. The results showed that the presence of silica colloids retarded the induction of gypsum crystallization and slightly decreased the nucleation rate. This was presumably due to the slightly increased critical size for nuclei as a result of the collision between silica colloids and newly formed nuclei. It was also found that gypsum crystallization rate was increased in the presence of silica colloids. This was attributed to the fact the silica colloids changed the morphology of gypsum crystals and resulted in larger crystals, as revealed by scanning electron microscopy (SEM) images. In the absence of silica colloids, there were mostly needle-shaped crystals along with a significant amount of clusters of 141 smaller crystals. On the contrary, when silica colloids were present, there were some needle-shaped crystals with much larger plate—shaped crystals. 142 5.1. Introduction Fouling - the undesired deposition and accumulation of materials on surfaces - is one of the most serious problems as it lowers the efficiency of many industrial processes. It was estimated that fouling-related costs correspond to ca. 0.25% of USA’s gross domestic product (GDP) [1]. In aqueous systems, fouling can be especially severe; membrane- based separation [2] (reverse osmosis and nanofiltration [3, 4], ultra- and microfiltration [5, 6], membrane distillation [7, 8], membrane bioreactors [9, 10]) and heat transfer [11] (such as tubular heat exchangers [12], and plate heat exchangers [13]) technologies suffer significant losses due to this phenomenon. Fouling reduces the quality and quantity of permeate, increases energy consumption, shortens equipment lifetime, and thus raises both operational and capital costs. Because of the complexity of natural waters, fouling is usually caused by a suite of distinct phenomena, each corresponding to a different fouling mechanism. For instance, in reverse osmosis (RO) membrane systems, precipitation of sparingly soluble salts (scaling) is accompanied by biological and particulate fouling. In heat exchange systems, scaling and particulate fouling are known to occur simultaneously [14]. While contribution of one type of fouling can be dramatic, there is growing evidence that interactions between different fouling mechanisms may play an important [9, 14, 15] or even dominant role [16] in determining the overall system performance. 143 Scaling, or colloidal fouling — the usually undesirable precipitation process that can severely impact permeate flux and rejection of membranes [l7] and performance of heat exchangers ([18, 19]) - has been shown to be influenced by surfactants [20], ionic species [21, 22], and natural organic matter (NOM) [23]. Surfactants have been shown to have an effect on gypsum crystallization in terms of induction time, growth efficiency, surface energy and crystal structure [20]. The effects varied with surfactants types. The cationic surfactant cetyltrimethylammonium bromide (CTAB) decreased the induction time, and increased the growth efficiency and surface energy, while the anionic surfactant had the contrary effects. The presence of other precipitating salts affected the structure and strength of gypsum crystals [21]. Polyelectrolytes have been shown to affect gypsum crystallization and the effect depended on the concentration, molecular weight, and composition of the polyelectrolytes [24]. Jiang et al. [25] have shown that the effect of borax on gypsum scaling relied on its concentration: borax expedited gypsum scaling at low concentration, but retarded it at high concentration. Studying gypsum crystallization in a nanofiltration (NF) membrane system filtrating agricultural drainage water, Le Gouellec and Elimelech [23] demonstrated that the presence of NOM resulted in shorter gypsum crystals. It was also found that a single aquatic humic acid could not model the effect of the complex NOM on gypsum crystallization. Although both scaling and colloidal fouling have been intensively studied independently in their respective fields, there is very limited research in the context that these two fouling mechanisms occur simultaneously [26]. All studies published to date can be roughly grouped into two categories: a) fundamental studies on the nucleation and growth 144 of crystals in the presence of colloids and b) practical studies concerning performance of a specific technology. With no research published on the effect of particles on colloidal fouling of membranes; most of the available knowledge comes from studies on the performance of heat exchangers. In heat transfer systems, the effect of particles on scaling has been demonstrated to be complex [27]. Bansal et a1. [19] found that the effect of suspended particles on scaling depended on the crystallizing ability of particles. Crystallizing CaSO4 particles greatly enhanced gypsum formation because they provided extra nucleation sites. On the contrary, noncrystallizing alumina particles reduced the gypsum formation. They believed that this effect was due to 1) particles reduced the strength of crystals deposited and thus resulted in a higher crystal removal rate, and 2) particles on crystal surface acted as a distorting agent. Andritsos and Karabelas [14] showed that CaCO3 precipitation rate was greatly enhanced by fine aragonite (CaCO; polymorph) particles, but was not affected by silica colloids. McGarvey and Turner [28] examined the interplay between CaCO3 precipitation and colloidal deposition. It was found although the rate of CaCO3 precipitation was not affected by the silt and hematite (Fe203) particles, the deposition of the particles was enhanced by CaCO3 precipitation. The morphology of gypsum crystals is very important in industry. As pointed out in [20], it is important for gypsum precipitation processes involving subsequent solid-liquid separation, such as removal of by-product gypsum in phosphoric acid manufacturing 145 process [29]. During the reuse of calcium sulfate dihydrate, it is preferable to form large plate-shaped crystals [30]. The objective of this work was to study the effect of colloids on the scaling of calcium sulfate dihydrate (gypsum) in a batch system. Model feed waters were prepared by suspending silica colloids in calcium chloride and sodium sulfate solutions of different degrees of supersaturation for gypsum. The effect of colloids on gypsum crystallization was examined in terms of induction time, rates of nucleation and crystallization, and morphology of gypsum crystals. 5.2. Background There are three steps in a crystallization process: achievement of supersaturation, formation of nuclei (i.e., nucleation), and their growth into larger crystals [31]. The degree of supersaturation in solution is represented by the gypsum saturation index (Sg) as: 2+ 2- 7 2+ [Ca 17 2- [SO I g Ksp ’ 146 where K Sp is the solubility product for CaSO4-2H20; [Ca2+] and [803-] are molar concentrations of calcium and sulfate ions in solution, respectively; 7C 2+ and 7802‘ a are the activity coefficients for calcium and sulfate ions, respectively; K for 5P CaSO4-2H20 in water at 20 °C has been shown to be 2.623E-4 [32] . Davis approximation can be used to estimate the activity coefficient as a function of ionic strength: I +x/7 log y = —(Az2 1 — 0.21), (5.2) where z is the ion charge; I is the ionic strength (M); A z 0.5 for aqueous solutions. Induction time is defined as the time that elapses between the moment the supersaturation condition is achieved and the moment crystals are detected in the solution. Induction time is known to be a function of saturation index, agitation intensity, temperature, presence of impurity, etc. [31, 33] and can be viewed as consisting of three contributions [31, 34]: find =tr +tn +tg2 (5.3) where t, is the relaxation time required for the system to achieve a quasi-steady-state distribution of molecular clusters; tn is the time required for formation of a stable 147 nucleus, which will grow into crystals instead of dissolving into the feed solution [35]. t g is the time for the nucleus to grow to a detectable size. As the definition of tg implies, the value of tind depends on the sensitivity of the technique used for detecting the change in the physical properties of the system due to the formation of crystals [33, 34]. Two types of techniques - optical and electrochemical — have been used to determine the induction time for gypsum crystallization. Optical techniques include monitoring of scattered and transmitted light [35], measurement of turbidity [20, 36-38], and direct naked-eye observation [39]. Electrochemical techniques include measuring the conductivity of the feed suspensions [40] or the concentration of calcium in the solution using calcium selective electrodes [36, 41, 42]. The technique- specific value of the induction time is usually defined operationally as the time for one of the above parameters to change to a certain extent because of the formation of nuclei and crystals in the solution [33, 34]. According to the classical homogeneous crystallization theory, the homogeneous nucleation rate (J s) can be estimated as [31, 34, 43]: AG ~ J S = F exp[— fig], (5.4) where F is a frequency constant known as the pre-exponential factor and has a theoretical value of 1030 nuclei/(cm3-s), which has been used for gypsum crystallization 148 [20]; AGcrm-ca] is critical free energy for nucleation; k is the Boltzmann constant, which has a value of R/N A where R is the gas constant (J/(mol-K)) and N A is the Avogadro’s number (mol'l); T is the absolute temperature (K). AGm-n-w, is related to surface energy ( 73) and the critical size( Ram-cal), the minimum radius required for a stable nucleus, as: 2 47:73 Rcrilical (5'5) AGcritical = 3 , According to the classical homogeneous crystallization theory, the relationship between the critical size (Rcritical) and the surface energy of the gypsum nuclei is governed by Gibbs-Thompson equation as [31, 43]: 273Vm Rcritical = mi (5.6) g where Vm is the molecular volume for gypsum (m3/mol). This equation implies that surface energy tend to increase with increasing size of nuclei. Combing Eqs. (5.4-5.6), we can get an expression for nucleation rate which has been used for gypsum crystallization [20, 35] as: 149 -16 3V2N J3 = Fexp ”:5 m ’42 . (5.7) 3(RT) (In S g) The surface energy (ys) can be calculated from the equation below [20, 34]: 1 firiViNAfte) logtmd = 8+ 3 2 3 , (5.8) T (logSg) (2.3R) where B is a dimensionless constant. It is worthy noticing that although Eqs. (5.4) and (5.5) are based on the assumption that the crystallization is ideally homogeneous (spontaneous) which is not common in most laboratory systems, they have been widely used for estimating the properties of gypsum crystals formed in stirred batch systems [20, 35, 37, 44, 45]. To make our results comparable to those studies, the two equations were adopted in this study to estimate the nucleation rate and surface energy of gypsum crystals. For bulk crystallization, the rate of gypsum crystallization can be expressed as [46]: dm —=k C -C , 5.9 dt b( b s) ( ) 150 where m is the amount of gypsum formed; kb is the gypsum bulk crystallization rate (also called diffusion-controlled crystallization constant [46]); Cb is the CaSO4 concentration in bulk solution; C S is the saturation concentration of CaSO4. 5.3. Material and methods 5.3.1. Experimental apparatus The experimental batch system is shown in Fig. 5.1. A 1,500 mL Pyrex® beaker with a 1,000 mL of feed suspension was placed in a water bath on the top of a multi-stirrer (model 1286, Lab-Line Instruments Inc., Melrose Park, IL). The temperature of the water in the bath was maintained at (20 :1: 0.2) °C using a programmable circulating chiller (model 9512, PolyScience, Niles, IL) equipped with an external resistance temperature detector. Keeping the feed temperature constant is very important because temperature influences not only gypsum crystallization but also measurement of conductivity. The water in the bath was mixed by two 3 inch magnetic stirrer bars at a rate larger than 200 rpm. The feed suspension in the beaker was mixed by an identical stirrer bar at a rate of 120 rpm. The suspension was continuously circulated by a gear pump (model EW—74013- 20, Cole-Partner, Vernon Hills, Illinois) at a flow rate of 1.6 L/min. A small portion of the flow was diverted into a flow-through UV-vis sample cell (model 178.710 — QS, Hellma GmbH & Co KG, Miillheim) for the on-line transmittance measurement using a spectrophotometer (Multi-Spec 1501, Shimadzu, Kyoto, Japan). The transmittance in the 151 Conductivity meter [1 Spectrophotometer _ii-‘l A“ . 9 Data acquisition system i- '(-' " ' E i l O : External temperature sensor Gear purrp ' ri> .................. Cooling liquid line I l E ..z : 20" . . It}? ‘ . OO— - Water bath and the multl-stirrer Chiller Figure 5.1. Schematic of the experimental batch system 152 range of 400 nm to 800 nm was recorded in real-time. The probe (model 013005A, Thermo Electron Corp. Beverly, MA) of a conductivity meter (model Orion 550, Thermo Electron Corp.) was placed in the suspension for on-line conductivity measurement. The data from the conductivity meter and the spectrophotometer were logged to a PC every 30 sec or 2 min. 5.3.2. Regents, silica colloids, and feed suspensions All reagents were of ACS analytical grade or higher purity (Fisher Scientific, Pittsburgh, PA) and were used without further purification. The deionized water used in the experiments was supplied by a commercial ultrapure water system (Lab Five, USFilter Corp., Hazel Park, MI) equipped with a terminal 0.2 pm capsule microfilter (PolyCap, Whatman Plc., Sanford, ME). The resistivity of the water was greater than 16 MQ-cm. No buffers were added to the feed water. Silica colloids (SnowTex-XL, Nissan Chemical America Corp., Houston, TX) were chosen as the model colloids. Light scattering data (ZetaPALS, Brookhaven Instrument Corp., Holtsville, NY) showed that the particle size distribution was relatively narrow (polydispersity factor of 0.074) with a mean hydrodynamic particle diameter of (55.6 $1.3)nm [47]. Silica colloids were chosen because of the abundance of the particulate silica in water. The experiments were conducted by adding CaC12-2H20 and Na2S04 separately. The amount of CaClz°2H20 and Nast4 added is shown in Table 5.1. 153 Table 5.1. Composition of model solutions Saturation index (S g) CaClz-2H20, g/L Na2804, g/L Ionic strength, mM 2.5 3.80 3.67 155.1 3.0 4.35 4.21 177.8 4.0 5.38 5.20 219.6 5.0 6.24 6.02 254.5 6.0 7.09 6.85 289.4 154 5.3.3. Characterization of the gypsum crystals At the point in the crystallization experiment that corresponded to the 70% decrease in transmittance, 10 mL of the feed suspension was taken from the feed vessel using a syringe and was filtered through a polycarbonate track-etch membrane (Nuclepore®, Whatman Plc.) with a 0.1 pm nominal pore size. Immediately after the filtration, 5 mL of methanol was passed through the membrane to remove the residual suspension on membrane surface to prevent formation of new gypsum crystals. The membrane was then dried in air at room temperature. A ca. 1 cm2 piece of the dried membrane was cut out and coated with osmium for 30 s using an osmium coater (Neoc-AN, Meiwa Shoji Co. Ltd., Kyoto, Japan). Scanning electron microscopy (SEM) images of the membrane surface were recorded using Hitachi S-47OOII Field Emission scanning electron microscope (Hitachi High Technologies America, Inc. Pleasanton, CA). The microscope was operated in the ultra-high resolution mode with an accelerating voltage of 15 kV and an emission current of 10 11A. The size distribution of gypsum crystals in the feed tank was measured using a particle Sizer (Malvem Mastersizer 2000, Malvem Instruments Ltd, Worcestershire, UK). When the transmittance of the feed suspension decreased by 70%, a 150 mL sample was taken from the feed vessel and placed into the Hydro SM® small volume sample dispersion module of the particle sizer. The initial feed suspensions were used as the background to remove the effect of ions and silica colloids. The sample in the dispersion module was stirred at a speed of 2,000 rpm. Data reported in this paper represent the average size 155 distribution computed based on the three data sets automatically recorded with a time interval of 5 s between them. 5.4. Results and discussion 5.4.1. Determination of induction time In this work, transmittance, calcium ion concentration, and conductivity were used and compared for determining the induction time. In our experiments, transmittance of the feed suspensions in the range of 400 nm to 800 nm was monitored in real-time. We used the 5% decline in transmittance as the starting time of gypsum crystallization as proposed by Lancia and co-workers’ [35]. We found that, a lower wavelength led to a relatively lower induction time (Fig.SS.I in Supplementary Data), which indicated that transmittance at a lower wavelength was more sensitive for detecting the change in the properties of feed suspension due to nucleation. For example, in the experiment with a supersaturated suspension (S g = 5) in the absence of silica colloids, the results showed that the induction time was 23 min, 27 min, and 29 min based on the transmittance at 400 nm, 600 nm, and 800 nm, respectively. The induction time reported in this work was based on the transmittance at 400 nm. A calcium selective electrode (Orion 97-20 Ionplus®, Thermo Electron Corp.) was also used to determine the induction time. The electrode measures the calcium potential difference across the sensing membrane between the target solution and the reference 156 -37 $ a 1 4 1.35 -38 s 1.3 E. 9 _ 1.25 3 e -39 e c 8 0 out 1.2 < 8 e 5 '40 1.15%: 0 - T. O o 1.1 3 -41 1.05 _42 c. ................................................................................... 1 Time (min) 100, .................. j ............ i ____________ . . .......... 13.8 .- . , (b) '§?«(((((“?«( 2 13 7 80 ' :3 13.6 g g 60 _ . E ,‘3 El Transmittance 13.5 < E - - 5 g 40 _ ........................................... K4 .................................. g i a 3 93513.4 :4 .3. 20 _ ............................................... ifs ................................. 13‘s 0 1 1 l 1 ...-T.mil”iililiiiliiiiiliiiilililiiiiiil’.‘ 132 0 10 20 30 40 50 60 70 Time (min) Figure 5.2. Comparison of different techniques for measuring gypsum induction time (S g = 5; no silica colloids): (a) calcium potential and conductivity measurement for determining the induction time; (measurements were for the diluted permeate of the UP membrane filtration); (b) transmittance and conductivity measurement for determining the induction time (measurements were for the feed suspension). 157 solution. The calcium potential is influenced by the ionic strength and temperature of the target solution. The measurement required an equilibration time of ca. 25 min, which was too long for in-situ measurement in our experiments. Thus, an experiment described below was conducted to determine the change of calcium ion concentration in the feed suspension. A gypsum crystallization experiment was carried out at a saturation index of 5 in the batch system without silica colloids. Every 5 minutes, a 5 mL suspension was taken from the feed suspension and filtered through an ultrafiltraion (UF) membrane using a stirred stainless steel dead-end filtration cell (HP 4750, Sterlitech Corp., Kent, WA) at a transmembrane pressure of 150 psi (1.03 MPa) to remove the possible nuclei formed in the feed suspension. The UF membrane (Omegam, Pall Corp., East Hills, NY) has a nominal molecular weight cut-off (MWCO) of 1,000 Da and a CaSO4 rejection of ca. 4.6% at this transmembrane pressure. A 2 ml permeate was collected and diluted 15 times with deionized water. The dilution ensured that 1) the volume was enough for the measurement, and 2) the diluted permeate solution was much less saturated, and no new nuclei formed during the measurement. Both the calcium potential and conductivity of the diluted permeate solution were measured. The results are shown in Fig. 5.2(a), from which it can be seen that neither calcium potential nor conductivity of the diluted permeate could clearly indicate the induction time. The above experiment was repeated with continuous measurement of the conductivity and transmittance of the feed suspension, as reported in Fig. 5.2(b). After 23 min, transmittance decreased by 5%. However, conductivity did not decrease until 38 min. This comparison indicates that, in our study, transmittance was more sensitive than 158 conductivity for determining the induction time. One possible reason for this is that, at the beginning of nucleation, light transmittance could detect the extremely small nuclei, which do not cause conductivity decline because they are still dissolved in the suspension. 5.4.2. Effect of silica colloids on the induction of gypsum crystallization The transmittance and conductivity of the feed suspensions as a function of time at different saturation indices were shown in Fig. 5.3. It was found that the presence of silica colloids resulted in a longer induction time. This effect was more evident at a low saturation index (F ig. $52 in Supplementary Data). The induction time with or without silica colloids at different saturation indices was shown in Fig. 5.4 (a). There is possibly a certain amount of chemicals, such as surfactants, in the stock suspension of the silica colloids. An eXperiment was conducted to test the effect of these chemicals on gypsum nucleation. First, 0.095 mL (the same amount as used in our experiments) of the stock silica suspension was taken and added into to 50 mL deionized water. Second, the diluted suspension was filtered through a 10 kDa ultrafiltration (UF) membrane to obtain a solution free of particles. The 10 kDa membrane was chosen to ensure a) the complete removal of silica particles and b) zero rejection of all chemicals present 159 1:] No colloids 80 _,_ 0 With colloids Ea é » . ., ‘_“.' 3 60 .. i E r: 40 i! '— N O 0 16 15 15 o "’ o 35 14 a = c 3 1:1 No Colloids 14 3 ° 13 . . 5. 2I 0 With coIIOIds .2 E 13 ’3‘ 2 12 91 in O 5, 12 '3 a E 11 11 g ‘5 1E 10 g 2. 8 10 a. 9 9 1 1000 100 Time (min) Figure 5.3. The transmittance (a) and conductivity (b) of the feed solutions as a function of time (transmittance was reported at a reduced frequency) 160 in the silica suspension. Third, a gypsum crystallization (S g = 4) experiment was conducted with the solution added into the feed suspension. The recorded induction time was compared to that in the experiment without adding the solution, and no evident difference was observed. This experiment indicates that the observed effect on gypsum crystallization was due to the colloids and not due to the chemicals in the stock suspension. It was also found that after conductivity started to decline, it followed a straight line 9Fig. 5.3). This was used for determining the gypsum crystallization rate, as discussed in Section 5.4.4. Based on the comparison of Fig. 5.3(a) and Fig. 5.3(b), it can be seen that transmittance measurement was more sensitive for measuring induction time than conductivity measurement, as also indicated in Fig. 5.2(b). Based on Eq. (5.6) and the surface energy, Ram-ca, for different saturation indices is estimated and shown in Table 5.2. We suspect that the increase in critical size in the presence of silica colloids in our study was due to the collision between the silica colloids and newly formed nuclei. The collision made newly formed nuclei more unstable. As a result, nuclei had to grow into a larger size to stay stable in the suspension instead of dissolving back into the suspension. 161 Table 5.2. Estimated critical size of nuclei in the absence and presence of silica colloids Sg Critical size of gypsum nuclei (Rcrmcal , nm) In the absence of colloids In the presence of colloids 3 0.609 0.628 4 0.482 0.496 5 0.416 0.427 6 0.373 0.383 162 This increase in Rem-”ca, retarded the induction, resulting in a larger induction time and a higher surface energy. 5.4.3. Effect of silica colloids on the surface energy and nucleation rate of gypsum crystallization Based on Eq. (5.8), log(t) was plotted against 1/(log S g )2 at different saturation indices (Fig. $53 in the Supplementary Data). From the slope, the surface energy was calculated to be 10.9 :t 1.2 J/(mol-K) ( p2 = 0.99) for the crystals formed in the absence of silica colloids and 11.2 i 1.1 J/(mol-K) ( p2 = 0.99) for those in the presence of 50 mg/L silica colloids. This indicates that the presence of silica colloids slightly increased the surface energy of crystals. Based on Eq. (5.7), the theoretical nucleation rates of gypsum crystallization at different saturation indices were calculated and presented in Fig. 5.4(b). The presence of silica colloids slightly decreased the nucleation rates. 163 \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ ,\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ E] N il Id With ll id 4 au 0 olloids on (a) induction time, (b) nu 22222 0 mo. 000000 o. 0. mosaic—... “$588 :23 22.2 r... .538. 2:3..mnbo rate, and (c) cleation te 5.4. Effect of silica c Figure crystallization ra 5.4.4. Effect of silica colloids on the gypsum crystallization rate In our experiments, the feed suspension volume is always 1 L, which enables us to change the Eq. (5.9) to: dC . gt-=kb(Cb-Cs), (5.10) where k'b is the crystallization rate (min'l), which has been shown to be dependent on the saturation index [48, 49]. Transmittance can not be used to determine the change in the concentrations because it depends on both the number and size of the nuclei and crystals [50]. In our experiments, the decline in conductivity after induction time was used to calculate the crystallization rate. The measurement was easy and straightforward. The decline of conductivity is proportional to the decrease of [Ca 2+] and [802'] , due to the formation of gypsum crystals. Thus the decline rate of conductivity (calculated from the slopes of the lines for conductivity versus time (Fig. 5.3)) can be used for determining dC/ dt based on a previously recorded standard curve of conductivity and CaSO4 concentration. Gypsum crystallization rates at different saturation indices in the absence and presence of silica colloids were shown in Fig. 5.4(c). At all of the saturation indices, the presence of silica colloids resulted in a higher gypsum crystallization rate, with the difference being larger for higher degrees of saturation. We suspect that this effect was due to that fact that 165 silica colloids influenced the epitaxial growth and morphology of gypsum crystals, which is discussed in Section 5.4.6. 5.4.5. Effect of silica colloids on crystal size distribution Laser diffraction-based analysis of gypsum crystal size distribution has been used by Mahmoud [20]. However, it should be noticed that the laser analyzer’s measurement is usually performed assuming the particles are spherical. Because the gypsum crystals are needle-shaped or plate-shaped, extreme care should be taken when one reports the size data based on this type of measurement. Thus, in this study, it was used only for qualitative analysis of gypsum crystal size distribution. Our measurement for crystal size distribution is shown in Fig. 5.5, which indicates there was a difference in the crystal size distribution in the absence and in the presence of the silica colloids. 5.4.6. Effect of silica colloids on the morphology of gypsum crystals Scanning electron microscopy (SEM) imaging was used for a direct observation of the morphology of the gypsum crystals. As shown in Section 5.3.3, when transmittance decreased by 70% at a saturation index of 5, 10 mL of feed suspensions was taken from the feed suspension. The samples were taken at 40 min and at 44 min for the experiments 166 No Colloids l l With Colloids M (.0 4:8 0'1 ._ _W a 1 l . l . .. o c _ a i 1__; TT4:_.__ > i ' i ' MHz? - lELL-11 U 1 ..--..1 10 103 1000 $00 Particle size (pm) Figure 5.5. Size distribution of gypsum crystals formed in the absence and in the presence of 50 mg/L silica colloids. Conditions: S g = 5, transmittance = 50% (Laser obscuration for the measurement without and with colloids is 9.7% and 11.3%, respectively.) 167 without and with silica colloids, respectively. This choice of sampling at the same transmittance ensured that the crystals had the same time periods of growth. After the samples were filtered through the 0.1 pm polycarbonate (PC) track-etch membrane, 5 mL of methanol was used to remove the residual water on the membrane surface, which prevented further growth of the crystals on membrane surface. This approach has increased the comparability of the morphology of crystals for identifying the effects of silica colloids. The SEM images in the absence and presence of silica colloids are shown in Fig. 5.6. 5.4.6.1. Length-to-width ratio From Fig. 5.6, it can be seen that, in the absence of silica colloids, most of the crystals were needle-shaped with a typical length-to-width ratio of ca. 20. In contrast, when the silica colloids were present, most of crystals were plate-shaped, with a typical length-to- width ratio of ca. 2. There were still some needle-shaped crystals, the structure of which was not uniform in terms of length-to-width ratio. There were a significant amount of clusters of smaller crystals (Fig. 5.5 (a-d)) observed in the silica colloid-free feeds. Having observed the needle-shaped crystals only at later stages of growth, we consider smaller crystals to correspond to earlier stages in the development of needle-shaped crystals. The clusters of smaller crystals were not found in feeds containing silica colloids. 168 5.4.6.2. Implications of epitaxial growth of gypsum for crystal morphology It was reported that the presence of cetyltrimethylamonium bromide (CTAB) slightly increased the surface energy of gypsum crystals, whose length-to-width ratio was smaller compared to those formed with no additives [20]. A similar result was observed elsewhere [51]: 0.1 g/L nitrilotrimethylenephosphonic acid (NTMP) increased the surface energy of gypsum crystals by ca. 50% and resulted in the formation of much larger plate- shaped crystals. Based on these studies, a tentative conclusion can be made that gypsum crystals with a higher surface energy tend to form larger and plate-shaped crystals, compared to those with a lower surface energy at the same conditions. In our experiments, however, the presence of silica colloids slightly increased the surface energy of crystals, as shown in Section 5.4.3. We suspect that the formation of large plate-shaped crystals in the presence of silica colloids was due to the effect of the colloids on the epitaxial growth of the gypsum crystals. As discussed in the references (e.g. [46] and [43]), the three key growth sites, kinks, steps, and terraces, dominate the crystal growth process. The size of the crystals depends on the growth rates of the three sites while the structure (shape) depends on the relative importance between the three growth rates. The presence of the silica colloids was expected to influence the mixing condition near the crystal surface [52]. We suspect that the resultant change in local mixing enhanced the surface flux of the crystals and 169 Figure 5.6. SEM image of gypsum crystals formed in the absence (a-d) and in the presence (e-h) of 50 mg/L silica colloids. Conditions: S g = 5, transmittance = 50%) 170 thus created more kinks and steps. This effect had two immediate consequences on crystal growth: 1) it increased the overall crystal growth rate, as we observed in conductivity measurement, and 2) it changed the epitaxial growth of the crystals and led to large, plate-shaped crystals. 5.5. Conclusions The effect of silica colloids on gypsum scaling was studied in a stirred batch system. Three techniques, light transmittance, conductivity and calcium ion concentration measurement, were compared for determining induction time. It was found that light transmittance was the most sensitive. The presence of silica colloids retarded the induction of gypsum crystallization and slightly decreased the nucleation rate. This was presumably due to the slightly increased critical size for nuclei, as a result of the collision between silica colloids and newly formed nuclei. The silica colloids increased the gypsum. crystallization rate. This was attributed to the fact that the silica colloids changed the morphology of gypsum crystals. In the absence of silica colloids, there were mostly needle-shaped crystals, along with a significant amount of clusters of smaller crystals. On the contrary, when silica colloids were present, there were fewer needle-shaped crystals with much larger plate-shaped crystals. 171 Acknowledgments Financial support of this work by the National Water Research Institute (project no. 05- TM-OO7) is gratefully acknowledged. We thank Pall Corp. for providing the UP membrane samples and Nissan Chemical America Corp. for supplying the silica suspension. We thank Ms. Ewa Danielewicz from the Center for Advanced Microscopy at Michigan State University for her assistance with osmium coating of SEM samples. 172 Nomenclature A dimensionless coefficient B dimensionless constant C gypsum concentration (mol/L) C b gypsum concentration in bulk solution (mol/L) C s saturation concentration of gypsum (mol/L) F frequency constant f (9) correction factor I ionic strength (M) J3 nucleation rate (nucleus/(cm3-s)) k Boltzmann constant (1 .38>< 10'23 UK) kb gypsum bulk crystallization rate (g'L/(min-mol» k}, gypsum crystallization rate (min'l) Ksp solubility product m mass of gypsum crystals (g) N A Avogadro’s number (mol'l) R gas constant (J/(mol-K)) Ram-ca] critical size for nuclei (m) S g saturation index T absolute temperature (K) tind induction time (min) V m molecular volume (m3/mol) 2 ion charge AGcrm-cal critical free energy for nucleation (J) Greek letters 7 activity coefficient 75 activity coefficient p2 correlation coefficient 173 Reference 10. ll. 12. S. J. Pugh, G. F. Hewitt and H. Muller-Steinhagen, Fouling during the use of seawater as coolant - the development of a user guide, Heat Transfer Engineering 26 (2005) 35 - 43. Committee Report: Recent advances and research needs in membrane fouling, J. Am. Water Works Assoc. 97 (2005) 79-89. R. Y. Ning, T. L. Troyer and R. S. Tominello, Chemical control of colloidal fouling of reverse osmosis systems, Desalination 172 (2005) 1. W. Song, V. Ravindran, B. E. Koel and M. Pirbazari, Nanofiltration of natural organic matter with H202/UV pretreatment: fouling mitigation and membrane surface characterization, J. Membr. Sci. 241 (2004) 143. S. Sethi and MR. Wiesner, Modeling of transient permeate flux in cross—flow membrane filtration incorporating multiple particle transport mechanisms, J. Membr. Sci. 136 (1997) 191-205. T. Carroll, S. King, S. R. Gray, B. A. Bolto and N. A. Booker, The fouling of microfiltration membranes by NOM after coagulation treatment, Wat. Res. 34 (2000) 2861. M. Gryta, M. Tomaszewska, J. Grzechulska and A. W. Morawski, Membrane distillation of NaCl solution containing natural organic matter, J. Membr. Sci. 181 (2001) 279. Surapit Srisurichan, Ratana Jiraratananon and A. G. Fane, Humic acid fouling in the membrane distillation process, Desalination 174 (2005) 63. R. S. Trussell, R. P. Merlo, S. W. Herrnanowicz and D. Jenkins, The effect of organic loading on process performance and membrane fouling in a submerged membrane bioreactor treating municipal wastewater, Wat. Res. 40 (2006) 2675. S. Rosenberger, C. Laabs, B. Lesjean, R. Gnirss, G. Amy, M. Jekel and J. C. Schrotter, Impact of colloidal and soluble organic material on membrane performance in membrane bioreactors for municipal wastewater treatment, Wat. Res. 40 (2006) 710. J. Taborek, T. Aoku, R. B. Ritter, Paeln J.W. and JG. Knudsen, Fouling - the Major Unresolved Problem in Heat Transfer, Chem. Eng. Prog. 68 (1972) 69 - 78. D. Butterworth, Design of shell-and-tube heat exchangers when the fouling depends on local temperature and velocity, Appl. Therm. Eng. 22 (2002) 789. 174 13. 14. 15. l6. 17. 18. 19. 20. 21. 22. 23. 24. 25. S. Grandgeorge, C. Jallut and B. Thonon, Particulate fouling of corrugated plate heat exchangers. Global kinetic and equilibrium studies, Chem. Eng. Sci. 53 (1998) 3050. N. Andritsos and A.J. Karabelas, The Influence of particulates on CaCO3 scale formation, J. Heat Transfer 121 (1999) 225-227. S. Lee, J. Cho and M. Elimelech, Combined influence of natural organic matter (NOM) and colloidal particles on nanofiltration membrane fouling, J. Membr. Sci. 262 (2005) 27—41. F. Wang and V.V. Tarabara. Coupled effects of scaling and colloidal deposition on the performance of reverse osmosis membranes. in The 233rd ACS National Meeting. March 2007. Chicago, IL. M. Uchymiak, A. Rahardianto, E. Lyster, J. Glater and Y. Cohen, A novel R0 ex situ scale observation detector (EXSOD) for mineral scale characterization and early detection, J. Membr. Sci. 291 (2007) 86. A. Karabelas, Scale formation in tubular heat exchangers--research priorities, International Journal of Thermal Sciences 41 (2002) 682. B. Bansal, H. Muller-Steinhagen and X. D. Chen, Effect of suspended particles on crystallization fouling in plate heat exchangers, J. of Heat Transfer, Transactions ASME 119 (1997) 568. M. H. H. Mahmoud, M. M. Rashad, I. A. Ibrahim and E. A. Abdel-Aal, Crystal modification of calcium sulfate dihydrate in the presence of some surface-active agents, J. Colloid Interface Sci. 270 (2004) 99-105. R. Sheikholeslami, Mixed salts -- scaling limits and propensity, Desalination 154 (2003) 117. Z. Amjad and J. Hooley, Influence of polyelectrolytes on the crystal growth of calcium sulfate dihydrate, J. Colloid Interface Sci. 111 (1986) 496. Yann A. Le Gouellec and Menachem Elimelech, Calcium sulfate (gypsum) scaling in nanofiltration of agricultural drainage water, J. Membr. Sci. 205 (2002) 279. M. Okazaki and S. Kimura, Scale formation on reverse-osmosis membranes, J. Chem. Eng. Jpn. 17 (1984) 145. W. Jiang, H. Pan, J. Tao, X. Xu and R. Tang, Dual roles of borax in kinetics of calcium sulfate dihydrate formation, Langmuir 23 (2007) 5070-5076. 175 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. R. Sheikholeslami, Composite fouling of heat transfer equipment in aqueous media - a review, Heat Transfer Engineering 21 (2000) 34 - 42. D. Hasson. Progress in precipitation fouling: a review, in:Understanding Heat Exchanger Fouling and its Mitigation, T.R. Bott et al. (Eds). 1999: Begell House, New York. GB. McGarvey and CW. Turner, Synergism models for heat exchanger fouling mitigation. In: T.R. Bott et al.Understanding Heat Exchanger Fouling and its Mitigation, Begell House, New York, (1999) 155-161. M. M. Rashad, H. M. Baioumy and E. A. Abdel-Aal, Structural and spectral studies on gypsum crystals under simulated conditions of phosphoric acid production with and without organic and inorganic additives, Cryst. Res. Technol. 38 (2003) 433 - 439. Y. Kojima and T. Yasue, Synthesis of large plate-like gypsum dihydrate from waste gypsum board, J. Euro. Ceram. Soc. 26 (2006) 777-783. J. W. Mullin, Crystallization. 4th ed. 2001, Oxford: Butterworth-Heinemann. D.K. Nordstrom, L.N. Plummer, D. Langmuir, E. Busenberg, H.M. May, B.F. Jones and D.L. Parkhurst, Revised chemical equilibrium data for major water- mineral reactions and their limitations, in Bassett, R.L. and Melchior, D. eds., Chemical modeling in aqueous systems 11: Washington DC, American Chemical Society Symposium Series 416, Chapt. 31, (1990) 398-413. 0. S6hnel and J. W. Mullin, A method for the determination of precipitation induction periods, J. Crystal Growth 44 (1978) 377 - 382. O. Schnel and J. W. Mullin, Interpretation of crystallization induction periods, J. Colloid Interface Sci. 123 (1988) 43-50. A. Lancia, D. Musmarra and M. Prisciandaro, Measuring induction period for calcium sulfate dihydrate precipitation, AIChE J. 45 (1999) 390 - 397. W.Y. Shih, K. Albrecht, J. Glater and Y. Cohen, A dual-probe approach for evaluation of gypsum crystallization in response to antiscalant treatment, Desalination 169 (2004) 213. S. He, J. E. Oddo and M. B. Tomson, The inhibition of gypsum and barite nucleation in NaCl brines at temperatures from 25 to 90[degree sign]C, Appl. Geochem. 9 (1994) 561. 176 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. D. Hasson, A. Drak and R. Semiat, Inception of CaSO4 scaling on RO membranes at various water recovery levels, Desalination 139 (2001) 73 - 81. W. H. Lam and K. M. Ng, Diastereomeric salt crystallization synthesis for chiral resolution of ibuprofen, AIChE J. 53 (2007) 429 - 437. P.G. Klepetsanis and PG. Koutsoukos, Kinetics of calcium sulfate formation in aqueous media: effect of organophosphorous compounds, J. Crystal Growth 193 (1998) 156 -163. Y.A. Le Gouellec and Elimelech. M., Control of calcium sulfate (gypsum) scale in nanofiltration of saline agricultural drainage water, Environ. Eng. Sci. (2002) 387-397. A. Hina, G. H. Nancollas and M. Grynpas, Surface induced constant composition crystal growth kinetics studies. The brushite—Gypsum system, J. Crystal Growth 223 (2001)213 -224. AS. Myerson, Handbook of Industrial Crystallization. 2nd ed. 2002, Boston: Butterworth-heinemann. 33-63. F. Alimi, H. Elfil and A. Gadri, Kinetics of the precipitation of calcium sulfate dihydrate in a desalination unit, Desalination 158 (2003) 9. T. A. Hoang, H. M. Ang and A. L. Rohl, Effects of temperature on the scaling of calcium sulphate in pipes, Powder Technol. 179 (2007) 31. G.H. Nancollas, The growth of crystals in solution, Adv. Colloid Interface Sci. 10 (1979) 215-252. P. Wang and V.V. Tarabara, Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance, J. Membr. Sci. 293 (2007) 111-123. B. R. Smith and F. Sweett, The crystallization of calcium sulfate dihydrate, J. Colloid Interface Sci. 37 (1971) 612-618. S. T. Liu and G. H. Nancollas, Linear crystallization and induction-period studies of the grth of calcium sulphate dihydrate crystals, Talanta 20 (1973). T. Naganurna and Y. Kagawa, Effect of particle size on light transmittance of glass particle dispersed epoxy matrix optical composites, Acta Mater. 47 (1999) 4321 - 4327. 177 51. 52. M. Prisciandaro, E. Olivieri, A. Lancia and D. Musmarra, Gypsum precipitation from an aqueous solution in the presence of nitrilotrimethylenephosphonic acid, Ind. Eng. Chem. Res. 45 (2006) 2070 - 2076. E. Meiburg, E. Wallner, A. Pagella, A. Riaz, C. Haertel and F. Necker, Vorticity dynamics of dilute two-way-coupled particle-laden mixing layers, J. Fluid Mech. 421 (2000) 185—227. 178 SS. SUPPLEMENTARY DATA 100 (((((((((((((((( ‘ :«o; :«(r r; , , fir, g 80 ._ ............................................... ........................... _ . Z A .............................................. _<>._: ............................ _ .\° * <2): 8 60 ' . '0'?) ........................... 3 g D Wavelength = 400 nm %% (2) g E O Wavelength=600 nm ----- [j-Cgs‘gw-W-f ------------ ~ «E, O Wavelength = 800 nm Cl 0 1 a: 40 , . . -. e 5 . - '- _ ............. a ........................................... 20 L ............................................. r ............................................ 0 L l A 0 10 20 30 Time (min) Figure 55.1. Transmittance as a function of time at three wavelengths (S g = 5; no particles) Fig. S5.1 shows that the induction time was 23 min, 27 min, and 29 min based on the transmittance at 400 nm, 600 nm, and 800 nm, respectively. The induction time reported in this work was based on the transmittance at 400 nm. 179 10.2 9 Asuaé 3.28 9 9.00 9. .................... ......................... S With colloids D No colloid O _ u ........................................ ............................. 0 0 6 4 3% 355538... b 0 8 2 300 400 Tlme (min) 200 100 Figure $5.2. Effect of particles on transmittance and conductivity (gypsum crystallization rate) at S g = 3, (a) Conductivity, (b) Transmittance 180 3 I I I I I I l E—y= 0. 49068 + o. 41353x R= 0.99292 ----- y= 0.48498 + o. 43729x R= 0.99655 2.5 I ........... . ......... _. -B— No colloids .0 2,0 --O--With colloids ‘ i ' ' c : T : :‘I 2 _ ..................................................... , .............................. J U? 2 1.5 —————— . -- ......... ........... .......... ......... _ 1 L 1 l 1 l 1 I 1 1.5 2 2.5 3 3.5 4 4.5 5 1l(logSg)2 Figure 85.3. The plots for calculating the surface energy of the crystals 181 CHAPTER SIX Wang, Fulin; Tarabara, Volodymyr V. (2008). Crystallization of gypsum in the presence of colloidal silica. II: in a reverse osmosis membrane system. (In preparation for publication in a journal) 182 CHAPTER SIX CRYSTALLIZATION OF GYPSUM IN THE PRESENCE OF COLLOIDAL SILICA. II: IN A REVERSE OSMOSIS MEMBRANE SYSTEM Abstract Gypsum scaling usually occurs along with colloidal fouling in reverse osmosis (R0) membrane systems treating brackish surface water and seawater, because pretreatment processes vary seasonally and can not guarantee a complete removal of colloidal particles in the R0 feed waters. This work presents the first study on the combined colloidal fouling and gypsum scaling in R0 membrane systems. The combined fouling was observed to have a significant synergistic effect on the performance of R0 membranes. In experiments with less saturated suspensions, gypsum scaling enhanced colloidal deposition and the combined fouling resulted in a much faster permeate flux decline, which was larger than the sum of the flux decline from colloidal fouling and that from gypsum scaling alone. In experiments with supersaturated suspensions, the presence of silica particles retarded induction time but increased the crystallization rate. When transmembrane pressure was applied, crystals that formed on the membrane surface through surface crystallization in the early stage greatly shortened the induction time and increased the rate of gypsum crystallization. 183 6.1. Introduction The decreasing water supply and the growing demand for water have led many water utilities to use alternative water sources, such as brackish surface water and seawater, with elevated levels of contaminants and salts [1, 2]. Meanwhile, the increasingly stringent water standards, such as Disinfectants/Disinfectants Byproducts Rule and Long Term 2 Enhanced Surface Water Treatment Rule, require a higher removal of contaminates to ensure human health [3]. These two factors have raised the popularity of reverse osmosis (R0) membrane filtration, which can serve as a reliable means for removing microorganisms, organics, precursors of disinfection by-products (DBPs), colloids, and other suspended matter of concern [3, 4]. However, membrane fouling, such as gypsum scaling and colloidal fouling, decreases the quality of product water, increases operating cost, and thus impedes a wider application of R0 membrane filtration. Scaling fi'om sparingly soluble salts, such as carbonate and sulfate, is a serious type of fouling for R0 membranes. Scaling occurs when the solubility of the salts is exceeded at a high recovery rate. Scales formed on a membrane surface reduce permeate flux, shorten membrane life and are difficult to be removed. Carbonate scaling can be controlled by decreasing the pH of the feed water. Calcium sulfate dihydrate (gypsum), however, is insensitive to pH, and can not be controlled by adjusting it. The effect of crossflow rate and system configuration on gypsum scaling in membrane filtration systems has been extensively examined. Lee and Lee [5] demonstrated that in 184 nanofiltration (NF) and R0 membrane systems, the gypsum scaling rate was decreased at a higher crossflow rate and a lower transmembrane pressure. Lee and Lueptow [6] used a rotating R0 membrane system to reduce the gypsum scaling. These studies achieved a reduction in gypsum scaling by decreasing the effect of concentration polarization, which resulted in a lower CaSO4 concentration at the membrane surface. Cohen’s research group [7] demonstrated a progressive axial development of gypsum crystals along the membrane surface, consistent with the increase in concentration polarization. Also, the membrane surface coverage by gypsum increased at a higher gypsum saturation index [7]. A substantial amount of suspended particles are present in seawater [8, 9], brackish water and wastewater [10, 11]. Pretreatment for R0 membrane systems varies seasonally [12- 14], and can not guarantee a complete removal of colloidal particles in the R0 feed waters [8, 9, 13, 15-17]. The colloidal particles can adhere to the inner pores of porous membranes and/or form a deposit layer on the membrane surface. The so-called colloidal fouling increases the hydraulic resistance to the permeate flow and thus decreases permeate flux. Numerous studies have been conducted on the effects of colloidal fouling on R0 membrane performance (e. g. [4] and the references therein). In practice, gypsum scaling usually occurs along with colloidal fouling in R0 systems treating brackish surface water or seawater. Although the pilot R0 membrane filtration system treating brackish the Colorado River water had a microfiltration (MF) membrane system as pretreatment, the R0 membrane was simultaneously fouled by colloidal fouling and scaling [18]. In another study, an R0 membrane filtration system in a 185 brackish water treatment plant used a pre-treatment process including coagulation, flocculation, dissolved air flotation and filtration (DAFF), pH correction, and cartridge filters [19]. However, an autopsy of the fouled R0 membrane clearly identified several kinds of foulants on the membrane surface, including colloids, extracellular polymeric substances (BPS), scales and natural organic matter (N 0M) [19]. Based on ca. 2 years of monitoring in a seawater R0 (SWRO) plant, Chua et al. [8] reported that although both a conventional pretreatment system and an ultrafiltration (UF) membrane system were used to remove the suspended particles [8], the silt density index (SDI) of the R0 feed water remained above 2.5 and a substantial amount of suspended particles remained. The SDI peaks indicated that the leakage of particles through membrane pretreatment was possible. Boerlage and co-workers [13] evaluated colloidal fouling in two R0 pilot desalination plants. In these two plants, the feed waters were pretreated by a slow sand filtration and a UF membrane, respectively, before entering the R0 membrane systems. It was found that both of the filtration methods were effective in removing particles larger than 0.45 pm, but not the smaller particles. Although there are numerous studies for the combined colloidal fouling and gypsum scaling in heat transfer systems (e.g. [20] and references therein), so far there has been no study available for the effects of combined colloidal fouling and gypsum scaling on the performance of R0 membranes. This study aimed at examining the interplay between the colloidal fouling and gypsum scaling in R0 membrane systems. A significant synergistic effect was observed during the combined gypsum scaling and colloidal fouling. 186 6.2. Experimental 6.2.1. Feed suspensions There are two pathways for gypsum scaling in membrane systems: surface crystallization and bulk crystallization [21-23]. In surface crystallization, gypsum crystals grow on the membrane surface. On the contrary, in bulk crystallization, gypsum nucleation occurs in the supersaturated feed solution and gypsum crystals deposit onto membrane surface. The rate of gypsum crystallization in membrane system can be expressed as [5, 24, 25]: dm E=k3(Cm—Cs)2, (6.1) for surface crystallization, and dm —=k C —C , 6.2 dt b( b s) ( ) for bulk crystallization, where m is the amount of gypsum formed, ks and kb are the gypsum surface and bulk crystallization rates (also called surface and diffusion controlled crystallization constant, [24], respectively), respectively, Cb and Cm are the CaSO4 concentrations in bulk solution and at the membrane surface, respectively, C s is the CaSO4 saturation concentration. Surface crystallization occurs at a low saturation index 187 on the membrane surface while bulk crystallization occurs at a high saturation index in the bulk feed water. The experiments were conducted by adding soluble CaC12-2H20 and Na2S04 separately instead of directly adding CaSO4°2H20. The advantages of this method included: (1) It was convenient to prepare supersaturated solutions with respect to calcium sulfate [26]; (2) The amount of undissolved salts was minimized [26]; (3) It was close to the situation in practice when the concentrations of ions are increased due to increase in recovery, the saturation index for CaSO4-2H20 exceeds 1, and gypsum formation begins; and, (4) Filtration of CaClz solution for 12 h conditioned membrane and reduced the effect of the change in the properties of membrane when in contact with salty suspensions. For experiments with less saturated feed suspensions, 35.9 g CaC12-2H20 and 34.7 g Na2S04 were dissolved separately and then were added into the deionized water. This resulted in an ionic strength of 0.073 M. Using the equations in our previous study [27], the saturation index (S g) of the feed suspensions was 0.91. In the control experiment with only soluble salts and silica particles, Na2804 was replaced by 28.5 g NaCl to achieve a close ionic strength and ion composition. 188 For experiments with supersaturated feed suspensions, 107.9 g CaC12-2H20 and 104.0 Na2804 were added, resulting in an ionic strength of 0.219 M and a saturation index of 4.04. In the control experiments, 85.7 g NaCl was used to replace N32S04. Feed channel spacers were used in these experiments. 6.2.2. Experimental apparatus The experimental apparatus has been described in our previous work [28]. The retentate flow rate was kept at 3.8 L/min. The 20 L feed suspension was mixed in a 30 L feed tank by the retentate flow and a 3 inches long magnetic stirrer bar. The verge of the membrane coupons were covered by impermeable transparent tape for 2 cm and 1 cm in width at the entrance and exit, respectively, to minimize the effect of inflow and outflow on gypsum scaling. The effective filtration area was reduced to 113.3 cm2 for each membrane coupon. After the experiments, one of the two membrane coupons was kept for microscopic analysis. Another one was put into stirred deionized water of 200 mL to dissolve the gypsum on the coupon surface. The concentration of calcium ion in the solution was measured using an atomic absorption spectrometer (AAnalysis 800, PerkinElmer Inc., Wellesley, MA) in the flame operation mode at a wavelength of 422.7 nm with a slit of 0.7 nm. Standard calcium solution was purchased from ULTRA Scientific Inc. (ECK-020, North Kingstown, RI). The standard curve had a correlation coefficient higher than 189 99.3%. The mass of gypsum on membrane coupon surface was calculated based on the measured concentration. 6.2.3. Reagents, silica particles and R0 membranes All reagents and chemicals were as reported previously [29]. Recently received silica particles (SnowTex-XL, Nissan Chemical America Corp., Houston, TX) were chosen as the model colloidal foulants. Polyamide thin-film composite R0 membrane (BW30-365, Dow-FilmTec, Minneapolis, MN) was used in all experiments. The characterization of the colloids and membrane, and the membrane storage procedure were as reported previously [29] as well. 6.2.4. Microscopical analysis The surface of fouled membrane samples was imaged using Hitachi S-47OOII F icld Emission scanning electron microscope (Hitachi High Technologies America, Inc. Pleasanton, CA) with a secondary electron detector. The microscope was operated in ultra-high resolution mode with an accelerating voltage of 15 kV and an emission current of 10 ILA. Samples were coated with gold for around 4 min at a current of 20 mA using a sputter coater (Emscope SC 500, Polaron Equipment Ltd. UK) before SEM imaging. The cross-section of the fouled membrane samples was imaged using Joel 6400V scanning electron microscope (Japan Electron Optics Laboratories, Japan) with a field 190 emission (Oxford EDS, Oxford Instruments, UK) in backscattered electron (BSE) imaging mode with an accelerating voltage of 20 kV. BSE imaging and X-ray energy dispersive spectroscopy (EDS) were used to examine structure of the fouling layer consisting of gypsum crystals and silica particles. We used following procedures to prepare the cross-section samples of fouled membranes for imaging. Samples were infiltrated with LR White resin (Electron Microscopy Sciences, Hatfield, PA) in gelatine capsules in agitation for 24 h. LR White resin is an organic chemical (80% Polyhydroxy substituted bisphenol A dimethacrylate resin and 19.6% C12 methacrylate ester) that does not dissolve gypsum crystals. Samples were polymerized at 60 °C for 24 h and were hand free cross-sectioned. Carbon was coated onto the cross-section with a carbon string evaporator (Ernest F. Fullam, Inc. Latham, New York.). 6.2.5. Protocols for membrane fouling experiments with less saturated suspensions Experimental protocol is similar to what we used in previous study [29]. After compacting membrane for 36 h and determining its hydraulic resistance, CaC12°2H20 was dissolved in the 20 L feed suspension to condition the membrane for 12 h. Na2S04 or NaCl was dissolved and added into the feed suspension. The initial permeate flux was controlled at 53 L/(mZ-h) to 54 L/(m2°h)(1.47'10‘5 m/s to 1.50405 m/s). In experiments with colloidal fouling, 1.90 mL ST-XL stock solution was added into the feed suspension to achieve a 50 mg/L silica particle loading. The pH of the feed suspension during the fouling experiments was between 5.8 and 6.2 and was not adjusted. The fouling experiment was stopped after 12 h. Membrane coupons were taken out and handled as 191 mentioned above. Afier each experiment, the crossflow system was cleaned thoroughly with a detergent solution. The pH of detergent solution was increased to 10 using NaOH solution. Then it was flushed with deionized water four times. The Teflon tubes in the retentate line were replaced by new ones to remove the possible silica particles deposited. To determine the mass of silica particles deposited on the membrane, a 25 mL sample was take from the suspension every 2 h and was diluted 2 times to dissolve the possible gypsum crystals. The absorbance at 220 nm of the resulted suspension was measured. The concentration of silica particles in the suspension was calculated based on the absorption using a standard curve recorded previously ( p2 = 1). A simple mass-balance equation was applied to calculate the mass of particles deposited on membrane surface (Mp). 6.2.6. Study the effect of silica particles on gypsum crystallization with supersaturated suspensions Four experiments were conducted to determine the effect of silica particles on gypsum crystallization with supersaturated suspension: (1) with silica particles, with transmembrane pressure; (2) without silica particles, with transmembrane pressure; (3) with silica particles, without transmembrane pressure; and (4) without silica particles, without transmembrane pressure. The transmittance at 400 nm of the suspension in the feed tank was continuously recorded using a spectrophotometer (Multi-Spec 1501, Shimadzu, Kyoto, Japan) by diverting a small portion of the retentate into a flow-through 192 UV—vis sample cell (Model 178.710 — QS, Helhna GmbH & Co KG, Miillheim) before returning this sample into the feed tank. A 5% drop in the transmittance was used to identify the beginning of induction. The reason for this choice was discussed in our previous study [27]. 6.3. Results and discussion 6.3.]. Combined colloidal fouling and gypsum scaling in less saturated suspensions 6.3.1.1. Effect on RO membrane performance Fig. 6.1 compares the permeate flux decline of the RO membrane during gypsum scaling, colloidal fouling, and combined gypsum scaling and colloidal fouling with less saturated suspensions, respectively. It can be seen that flux decline due to gypsum scaling after 12 h filtration was ca. 12%, close to that when the RO membrane was fouled by colloidal fouling alone. When gypsum scaling and colloidal particles were both present, the flux decline was ca. 42%, much larger than the sum of the decline from gypsum scaling and colloidal fouling alone. We suspect that three factors contributed to this observation. First, the silica particles occupied the open space between the crystals (as shown in Figs. 6.4-6.6). The resultant fouling layer consisted of large gypsum crystals (>20 pm in length and 2~3 pm in diameter) and tiny particles (55 nm in diameter). Because the gypsum crystals were not 193 “I I l . ' o I l I I ... I" k(t' ( ( - : ( (““W'R'It'm‘ “(£32. ................. "( ””‘lm ., 0_9 _ ............................... (flu X 3 _ '0: . +3 ~ s .. 0.8 — ........ E . 8 . . 1: ~ 3 3 . g 07 _. ........... . To “ 1 5 3 E 2 =. 1 z 0.6 D CaCl2 + Na2804 only 0 CaCl2 + NaCI + Particles 5 A CaCl2 + Na2$04+ Particles 3 0_5m1.l...l...l... 0 2 4 6 Filtration time (h) Figure 6.1. Comparison of membrane flux decline in the RO membrane system during gypsum scaling (indicated as “CaC12+ Na2SO4 only”), colloidal fouling (“CaC12+ NaCl + particles”), and combined fouling experiments (“CaC12+ Na2S04 + particles”). No spacers were used; ST-XL particles loading (if added): 50 mg/L; Sg (for gypsum scaling): 0.91; Initial hydraulic resistance of the membrane to pure water Rmo: (10.1 :1: 0.1)><10l3 m'l; Transmembrane pressure AP: (310 :i: 4)psi (2.37 MPa); Initial permeate flux Jo: (53.6 a: 0.5) L/(mz'h) (1.49-10's m/s); Retentate flow rate: 3.8 L/min; Feed suspension temperature: (20.0 :1: 0.5) °C. 194 permeable to water, the water passed the fouling layer through the open spaces between the particles or those between the particles and crystals. The effect of the crystals on the flow included larger tortuosity and “area obstruction” [30], which greatly increased the hydraulic resistance of the fouling layer. A theoretical model was developed in another study to predict the permeability of this type of porous media [31]. Second, gypsum scaling enhanced the deposition of silica particles, as shown in Fig. 6.3 and discussed later. Third, the retentate flow was less effective in removing the fouling layer resulted from the combined fouling than that from either gypsum scaling or colloidal fouling alone. After each experiment, a small portion of the fouled membrane was cut out and gently washed by deionized water from a wash bottle. SEM imaging of these washed samples demonstrated that after the cleaning, the fouling layer from gypsum scaling or colloidal fouling alone was effectively removed. On the contrary, on the surface of the membrane fouled by the combined fouling, a significant amount of crystals and particles remained after washing. The above three factors resulted in a much faster flux decline when the combined fouling occurred. Fig. 6.2 demonstrates the permeate conductivity of the RO membrane during gypsum scaling, colloidal fouling, and combined gypsum scaling and colloidal fouling, respectively. In these experiments, there were three or four ions in the permeate, all of which contributed to permeate conductivity. The RO membrane used in our study (BW30-365, Dow-FilmTec) had a rejection of higher than 99.5% for CaSO4 or CaClz and a NaCI rejection of 97% to 98%, which indicated that NaCl was the major contributor to conductivity. Thus, although it was not feasible to determine the concentration of CaSO4 195 Permeate conductivity (uslcm) b I l -1 Is . ( . . 100 6,1440%)", . . .' 33:5, ......... ‘. ....... LL11!” .............. 3 .............. ;. ............ _n e ’1”), [’5’]; I! Fflb’L-L _ ' ' . "IIIIHllllumm I" " "'lllliiII-I- E 80 -"- "'""!".'I_'Inin; ------- O CaCl2 + NaCl + Particles 9‘1"“? 'mmmnn A CaCl2 + Na2$04 + Particles Cl CaCl2 + Na2$04 only 60 ____________________ i. Filtration time (h) Figure 6.2. Permeate conductivity of the RO membrane during gypsum scaling, colloidal fouling, and combined fouling experiments, respectively. (Experimental conditions were shown in the caption of Fig. 6.1.) 196 and NaCl in permeate, respectively, based on permeate conductivity, it was acceptable to use the permeate conductivity as an “indictor” of salt rejection. As shown in Fig. 6.2, in the experiment with gypsum scaling only, the permeate conductivity increased for a short period and then decreased afterwards. This was in accordance with the observations by other researchers using CaClz and Na2SO4 for gypsum scaling experiments [32]. One possible reason for the short increase in permeate conductivity in the beginning is that initially some part of the ions on the membrane surface formed nuclei or crystals instead of passing through the membrane, and thus the concentration of salt at the membrane surface (Cm) decreased because of the formation of gypsum crystals on the membrane surface. In the experiment with colloidal fouling alone, the permeate conductivity increased because the fouling layer enhanced the salt concentration polarization, as discussed in [28, 33]. In the experiment with the combined gypsum scaling and colloidal fouling, the permeate conductivity decreased for the initial 3 h and increased afierwards. This indicated that after 3 h, the effect of enhanced salt concentration polarization by colloidal fouling was more important. 197 6.3.1.2. Effect of colloidal fouling on gypsum scaling As shown in Section 6.2, the mass of gypsum crystals on the membrane at the end of the experiments was determined by dissolving the crystals on the membrane surface in deionized water and measuring calcium ion concentration. The results showed that (246.1 :1: 4.7) mg gypsum formed in the experiment with gypsum scaling alone, and (242.9 :E 4.5) mg gypsum formed with the combined gypsum scaling and colloidal fouling. This indicated that the presence of the silica particles did not have evident effect on the amount of gypsum crystals on membrane surface. It is worthy noticing that, compared with the flux with gypsum scaling alone, the flux in the experiments with the combined fouling was significantly lower (Fig. 6.1), which was expected to result in a lower concentration of the ions on the membrane surface, according to the theory of concentration polarization. Thus less gypsum crystals were expected to form on membrane surface based on the kinetics of gypsum crystallization. However, as we indicated above, in the presence of silica particles the retentate flow became less effective in removing the fouling layer consisting of silica particles and gypsum crystals. As a result, gypsum crystals in the combined fouling experiment were more likely to stay on the membrane than those in the experiment with gypsum scaling alone. We suspect that this is the reason why we did not observe difference on the amount of gypsum crystals on the membrane surface in absence and presence of silica particles. It should be noted that the amount of calcium ion for the gypsum formed on the 198 350”T”?""!""I""I~~I~HI.. 300? _________ _: 250:— ----------- ........ E ____________ 3 ____________________________________ In“: 200_ j ...... M, _______ _: l- : I o r I— n ... u . o I 150 _ ................................... o ............................................ .2 l I I I I l- I e . n u .4 I I I I I I. I l I e e I I l 0 0 100 1 ---------------------- ' ' ' ‘ ' —-a— CaCl2 + Na2304 + Particles .-' , --O-- CaCl2 + NaCl + Particles )- . n j— ------------- . ......... 'u—.--n-a-c-I'fi'I'CCIFUIC‘III'I. IIIIIIIIIIIII 1f‘T IIIIIIIIII "v‘. n a u n u I. I I i c n r e v o I i l o o I I I Mass of particles deposited (mg) 0 2 4 6 8 10 12 Filtration time (h) Figure 6.3. Mass of silica particles deposited in the absence and presence of gypsum scaling (Experimental conditions were shown in the caption of Fig. 6.1.) 199 membrane surface accounted for less than 0.6% of calcium in the feed suspensions, which indicated that the amount of gypsum formed on membrane surface did not significantly affect the CaSO4 concentration in feed suspensions. 6.3.1.3. Effect of gypsum scaling on colloidal fouling Fig. 6.3 shows the amount of silica particles deposited on the membrane surface in the absence and presence of gypsum scaling. Although the permeate flux was lower (shown in Fig. 6.1), in presence of gypsum scaling more silica particles deposited on the membrane surface than in the experiment without gypsum scaling. We suspect that this was due to two factors. First, the gypsum crystals resulted in a rougher membrane surface. Colloidal particles were more likely to attach on a rougher membrane surface [34]. Second, as discussed above, the fouling layer consisting of both particles and gypsum crystals was more difficult to be removed by the retentate flow. As a result, in the presence of gypsum scaling, more silica particles deposited on the membrane surface. 6.3.1.4. Microscopical analysis of the fouling layer on membrane surface Figs. 6.4 and 6.5 show the SEM images of the crystals in the absence and presence of the silica particles, respectively. Without the silica particles, gypsum crystals grew from different spots and formed large and dendritic clusters of crystals. This is the typical structure of crystals formed through surface crystallization. In the presence of silica 200 Figure 6.4. Gypsum crystals in the absence of silica particles (Experimental conditions were shown in the caption of Fig. 6.1.) Figure 6.5. Gypsum crystals in the presence of silica particles at different magnifications (Experimental conditions were shown in the caption of Fig. 6.1.) 201 / Fouling layer \ ‘ ,.;¢ ' Membrane /' Figure 6.6. X-ray energy dispersive spectroscopy (EDS) analysis of the fouling layer consisting of gypsum crystals and silica particles (Experimental conditions were shown in the caption of Fig. 6.1.) 202 particles, the structure of the crystals did not change. The open space among crystals was occupied by the silica particles. It was also observed that in the absence of the silica particles, there were no crystals in the entry area. More crystals grew along the flow direction with most crystals appearing in the exit area. This to was due to increase in concentration polarization along the flow direction and the flush by the retentate flow, which was in accordance with that reported by other researchers [7, 35]. On the contrary, in the presence of silica particles, SEM imaging (pictures not shown) revealed that there were significant gypsum crystals in both the entry and exit areas. We suspect that this is also due to the fact the fouling layer consisting of silica particles and crystals was more difficult to be removed than that of crystals only. Fig. 6.6 reveals the structure of the fouling layer consisting of gypsum crystals and silica particles based on the X-ray energy dispersive spectroscopy (EDS) analysis of the cross- section of the fouling layer. It also demonstrates that the silica particles were embedded into the inclusion provided by gypsum crystals. The distribution of silica particles in the fouling layer was not uniform. Relatively more silica particles deposited closer to the membrane surface. 6.3.2. Combined colloidal fouling and gypsum scaling in supersaturated suspensions 6.3.2.1. Effect of silica particles on gypsum crystallization Fig. 6.7 shows the influence of silica particles on gypsum crystallization at different operating conditions with supersaturated suspensions in terms of transmittance and 203 conductivity. From the transmittance (Fig. 6.7(a)), it can be clearly seen that the presence of silica particles resulted in a slower transmittance decline and increased the induction time. This is in accordance with our previous study in a batch crystallization system [27]. It is also shown that when transmembrane pressure was applied, the induction time was greatly shortened. It is due to the fact that when transmembrane pressure was applied, the permeate flow increased the concentration on the membrane surface although the spacers were used in the feed channel. This led to some crystals forming on the membrane surface through surface crystallization (as shown in Figs. 6.9 and 6.10). These crystals acted as “seeds” and greatly decreased the induction time. As proposed in our previous study [27], the crystallization rate can be estimated based on the decline of conductivity. Eq. (6.2) can be written as, dC . E— = kb(Cb — C, ), (6.3) where k'b is the crystallization rate (min'l). Based on the conductivity data in Fig. 6.7(a), the crystallization rate under different operating conductions was estimated as: (1) Without silica particles, without transmembrane pressure: k1 = 0.00053 mini; (2) With silica particles, without transmembrane pressure: k2 = 0.00072 min"; (3) Without silica particles, with transmembrane pressure: k3 = 0.0011 min"; and (4) With silica particles, with transmembrane pressure: k4 = 0.0015 min". 204 0 With A P, No Part. "" A With A P, With Part. ) 3 g D No A P, No Part. 80 — <> NoAP,Wlth Part. ‘3 ~ (a) I ' ‘ ' o\ v P 8 60 _ ........................... 5 . 5 . g 40 _ ................................... h .— .— 20 _ ......................................................... ’- in}: L - o 1 l l I i l l l 1 l 1 l 12 ‘7 Y I I ! 1' I I Y F I l I I F r Y I T ! 1' T l V I V l T r I ! V T Y r ’1 II.- 4”!”,’” /: : I g ;;:-2.:z1;29;992w;*29fl“-" 1 i \’ b"i.'jj"h’\‘.rs.,”/\\%\ i /— - — 9-9/9. 9; : 118 _W' ........... ; ......... /// _./_I,,/. "an. "MM/9,, ........ .- ......... _ ' . . ///’ t ‘ . "" ,1: r-_\ .’ - . ' I . ?( - ' in”, / I A (b) I ///l Q?" Int "Q? . .1 g z - ..., 9,, s . a- . .. I J 0 1 ’////, a; " f) B ' 5 ’9. 3 V9,, ._ .............................................. ’ ....a.‘ ....... .......... WEI-”I, s11 6 . . . g . ’ 2 ff 1 I"‘ 5| 5‘ 0 With A P, No Part. A, 5 i9, 3 ’35:” 3? A With A P, With Part. "a; 9 3114 D NOAP,NO Part. _______ ’IA ________ ___________________ 'g o No A P, With Part. ; - j o . . . i 9 0 ' ' s ’4, ‘ - ' A A 11.2 _ .......................................................... A ......................... _ 1 1 1 l L L A 1 l l A l l A l l i l l 1 i 1 1 l l I: J l l 1 j l L L 1 0 50 100 150 200 250 300 350 400 Time (min) Figure 6.7. Effect of silica particles on gypsum crystallization at different operating conditions: (a) Transmittance of the feed suspensions; (b) Conductivity of the feed suspensions. (Silica particles (if added) loading: 50 mg/L; Transmembrane pressure (if applied): 300 psi; Experiments were conducted with spacers using supersaturated suspensions with S g = 4.04 .) 205 7’] 79",“ (:14! “— r r r r l r r ‘— 1 -””’/II//,),')“ .,.-—'».(.(««((‘<««««(«((«t(«t««««(«(«««««««««(««(««(((««««««t(«(««r- ”II,;"'='a‘-.=.- . - i 0'8 i : "‘ """ """""" ; """""" ’- """""" """"" --------- fl X : : : : : : ; . 3 a: _ 2 s s 2 s s s « 3 0.6 ....... .......... ............ ........... ......... _ g 2 I 1 i I ‘ g. ; : : : 5 i 3 o 4 5 5 5 ‘ .........EEWKQEQPPFPP ..... _ .5 ' 3 3 3 = meastrred in-i-llne . .-. z 2 £ , / a i J g i 5:515:- : : : g 0.2 ; 4 , ”Ix/Illl/l/l/I/I/I/I/I/l/l/l/H/l/Ill/[Illl/I/l/l/I'A‘ o CaCl2 + NaCl + Particles E ' D CaCl2 + Na2$04 only 0 A CaCl2 + Na2$04 + [Particles ’ I i 0 0.5 1 1.5 2 2.5 3 3.5 4 Filtration time (it). Figure 6.8. Normalized permeate flux at three experimental conditions. Experimental conditions were shown in the caption of Fig. 6.7.) 206 As shown in Fig. 6.7(b), the presence of silica particles increased the crystallization rate. This is in accordance with what we observed in the batch crystallization system [27]. Similarly, when the transmembrane pressure was applied, the crystals formed on the membrane surface greatly increased the crystallization rate. Bansal et al. [26] found adding gypsum crystals greatly enhanced gypsum formation because they provided extra nucleation sites. 6.3.2.2. Effect on RO membrane performance Fig. 6.8 compares the permeate flux decline of the RO membrane during gypsum scaling, colloidal fouling, and combined gypsum scaling and colloidal fouling with supersaturated suspensions, respectively. It can be seen that when scaling occurred in supersaturated suspensions, the permeate flux decreased dramatically after only ca. 2 h filtration, and the presence of silica particles led to a slightly faster permeate flux decline. 6.3.2.3. Microscopical analysis of the fouling layer on membrane surface Fig. 6.9 shows the crystals formed in the initial stage and after a 12 h filtration, respectively. It is found that in the initial stage the crystals were large and dendritic, which confirmed our statement that at the early stages crystals formed first on membrane surface through surface crystallization. Different to those shown in Fig. 6.4, the clusters in Fig. 6.9(a) had some large plate-shaped crystals. As shown in Fig. 6.9(b), after 12 h, the crystals were 207 Figure 6.9. The crystals formed in the absence of silica particles. (a) 20 min; (b) 12 h (Experimental conditions were shown in the caption of Fig. 6.7.) 208 '—s_ .‘iliiiiiii I i 12 3mm x1 m» SEiMi ill/5;“ )5 15 02 El'.‘ Mm Figure 6.10. The gypsum crystals formed in the presence of silica particles. (a) 20 min; (b)12 h. (Experimental conditions were shown in the caption of Fig. 6.7.) 209 mainly needle-shaped, which is the typical structure of gypsum crystals formed through bulk crystallization. Fig. 6.10 shows that the crystals formed in the presence of silica particles in the initial stage were also through surface crystallization. After a 12 h filtration, the crystals formed through bulk crystallization were mainly needle-like, similar to those formed in the absence of the particles. 6.4. Conclusions The interactions between colloidal fouling and gypsum scaling in RO membrane systems were investigated with both less saturated and supersaturated suspensions. A significant synergistic effect was observed during the combined colloidal fouling and gypsum scaling. When gypsum scaling was dominated by surface crystallization, gypsum scaling enhanced colloidal deposition and the combined fouling resulted in a much faster permeate flux decline, which was larger than the sum of the flux decline from colloidal deposition and that from gypsum scaling alone. When bulk crystallization dominated, the presence of silica particles retarded induction time, but increased the crystallization rate. When transmembrane pressure was applied, crystals formed on the membrane surface through surface crystallization in the early stage greatly shortened the induction time and increased the gypsum crystallization rate. This study implied that it was important to take into consideration the effects of combined fouling when designing and Operating RO membrane systems. 210 Acknowledgments Financial support of this work by the National Water Research Institute (project no. 05- TM-007) is gratefully acknowledged. We thank Dow-FilmTec for providing us with membrane samples and Nissan Chemical America Corp. for supplying the silica suspension. We also thank Dr. Ewa Danielewicz and Dr. Alicia Pastor from the Center for Advanced Microscopy at Michigan State University for their assistance with the cross-section sample preparation and EDS analysis. 211 Reference 10. 11. 12. 13. D. Bursill, Drinking water treatment - understanding the processes and meeting the challenges, Wat. Sci. Technol.: Water Supply 1 (2001) 1-7. State of California, Findings and Recommendations, (2003). United States Environmental Protection Agency, Report: Membrane filtration guidance manual (815-R-06-009), (2005). AWWA Membrane Technology Research Committee, Committee report: Membrane processes, J. Am. Water Works Assoc. 90 (1998) 91-105. S. Lee and CH. Lee, Effect of operating conditions on CaSO4 scale formation mechanism in nanofiltration for water softening, Water Res. 34 (2000) 3854. S. Lee and RM. Lueptow, Control of scale formation in reverse osmosis by membrane rotation, Desalination 155 (2003) 131. A. Rahardianto, W.Y. Shih, R.W. Lee and Y. Cohen, Diagnostic characterization of gypsum scale formation and control in RO membrane desalination of brackish water, J. Membr. Sci. 279 (2006) 655. K. T. Chua, M. N. A. Hawlader and A. Malek, Pretreatment of seawater: Results of pilot trials in Singapore, Desalination 159 (2003) 225-243. Harvey Winters, Twenty years experience in seawater reverse osmosis and how chemicals in pretreatment affect fouling of membranes, Desalination 110 (1997) 93. B. A. Winfield, A study of the factors affecting the rate of fouling of reverse osmosis membranes treating secondary sewage effluents, Water Res. 13 (1979) 565. Seung-Hyun Kim, Seong-Yong Moon, Cho-Hee Yoon, Seong-Keun Yim and Jae- Weon Cho, Role of coagulation in membrane filtration of wastewater for reuse, Desalination 173 (2005) 301. J. van Agtmaal, H. Huiting, P. A. de Boks and L. L. M. J. Paping, Four years of practical experience with an Integrated Membrane System (IMS) treating estuary water, Desalination 205 (2007) 26. S. F. E. Boerlage, M. Kennedy, M. P. Aniye and J. C. Schippers, Applications of the MFI-UF to measure and predict particulate fouling in R0 systems, J. Membr. Sci. (2003) 97-116. 212 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. A. D. Khawaji, I. K. Kutubkhanah and J. Wie, A 13.3 MGD seawater RO desalination plant for Yanbu Industrial City, Desalination 203 (2007) 176. M. Luo and Z. Wang, Complex fouling and cleaning in-place of a reverse osmosis desalination system, Desalination 141 (2001) 15-22. Y. R. Ning, L. T. Troyer and S. R. Tominello, Chemical control of colloidal fouling of reverse osmosis systems, Desalination 172 (2005) 1-6. Christopher J. Gabelich, Tae I. Yun, Bradley M. Coffey and I. H. Mel Suffet, Pilot-scale testing of reverse osmosis using conventional treatment and microfiltration, Desalination 154 (2003) 207. C. J. Gabelich, M. D. Williams, A. Rahardianto, J. C. Franklin and Y. Cohen, High-recovery reverse osmosis desalination using intermediate chemical demineralization, J. Membr. Sci. 301 (2007)l31. T. Tran, B. Bolto, S. Gray, M. Hoang and E. Ostarcevic, An autopsy study of a fouled reverse osmosis membrane element used in a brackish water treatment plant, Water Res. 41 (2007) 3915-3923. D. Hasson. Progress in precipitation fouling: a review, in:Understanding Heat Exchanger Fouling and its Mitigation, T.R. Bott et al. (Eds). 1999: Begell House, New York. A.G. Pervov, Scale formation prognosis and cleaning procedure schedules in reverse osmosis systems operation, Desalination 83 (1991) 77. M. Okazaki and S. Kimura, Scale formation on reverse-osmosis membranes, J. Chem. Eng. Jpn. 17 (1984) 145. S. Lee, J. Kim and CH. Lee, Analysis of CaSO4 scale formation mechanism in various nanofiltration modules, J. Membr. Sci. 163 (1999) 63. G.H. Nancollas, The growth of crystals in solution, Adv. Colloid Interface Sci. 10 (1979) 215-252. W.Y. Shih, A. Rahardianto, R.W. Lee and Y. Cohen, Morphometric characterization of calcium sulfate dihydrate (gypsum) scale on reverse osmosis membranes, J. Membr. Sci. 252 (2005) 253-263. B. Bansal, H. Muller-Steinhagen and X. D. Chen, Effect of suspended particles on crystallization fouling in plate heat exchangers, J. of Heat Transfer, Transactions ASME 119 (1997) 568. 213 27. 28. 29. 30. 31. 32. 33. 34. 35. F. Wang and V. V. Tarabara, Crystallization of gypsum in the presence of colloidal silica. I: in a stirred batch system, In preparation (2008). F. Wang and V.V. Tarabara, Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance, J. Membr. Sci. 293 (2007) 111-123. F. Wang and V.V. Tarabara, Coupled effects of colloidal deposition and salt concentration polarization on reverse osmosis membrane performance, J. Membr. Sci. (submitted) (2006). C. R. Ethier, Flow through mixed fibrous porous materials, AIChE J. 37 (1991) 1227-1236. F. Wang and V. V. Tarabara, Modeling the permeability of fiber-filled porous media using Kozeny-Carman-Ethier equation, In preparation (2008). M. Uchymiak, A. Rahardianto, E. Lyster, J. Glater and Y. Cohen, A novel R0 ex situ scale observation detector (EXSOD) for mineral scale characterization and early detection, J. Membr. Sci. 291 (2007) 86. E. M. Hoek and M. Elimelech, Cake-enhanced concentration polarization: a new fouling mechanism for salt-rejecting membranes, Environ. Sci. Technol. 17 (2003) 5581-5588. X. Zhu and M. Elimelech, Colloidal fouling of reverse osmosis membranes: measurements and fouling mechanisms, Environ. Sci. Technol. 31 (1997) 3654- 3662. A. P. Mairal, A. R. Greenberg, W. B. Krantz and L. J. Bond, Real-time measurement of inorganic fouling of RO desalination membranes using ultrasonic time-domain reflectometry, J. Membr. Sci. 159 (1999) 185-196. 214 .. m IIiflil'ijifliiilfliziflijieilIiflilieilijliil