PROTEIN TRANSPORT AND FOULING BEHAVIOR OF …

107
The Pennsylvania State University The Graduate School College of Engineering PROTEIN TRANSPORT AND FOULING BEHAVIOR OF ZWITTERIONIC ULTRAFILTRATION MEMBRANES A Thesis in Chemical Engineering by Mahsa Hadidi © 2014 Mahsa Hadidi Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science May 2014

Transcript of PROTEIN TRANSPORT AND FOULING BEHAVIOR OF …

The Pennsylvania State University

The Graduate School

College of Engineering

PROTEIN TRANSPORT AND FOULING BEHAVIOR OF ZWITTERIONIC

ULTRAFILTRATION MEMBRANES

A Thesis in

Chemical Engineering

by

Mahsa Hadidi

© 2014 Mahsa Hadidi

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Master of Science

May 2014

ii

The thesis of Mahsa Hadidi was reviewed and approved* by the following:

Andrew L. Zydney

Head of the Department of Chemical Engineering

Walter L. Robb Chair and Professor of Chemical Engineering

Thesis Advisor

Ali Borhan

Professor of Chemical Engineering

Themis Matsoukas

Professor of Chemical Engineering

*Signatures are on file in the Graduate School.

iii

ABSTRACT

The need for robust, cost-effective, and high-resolution methods for the

purification of recombinant proteins has created a need for ultrafiltration membranes with

lower fouling behavior and high selectivity. Several recent studies have shown that

zwitterionic membranes can be low fouling while retaining some of the benefits of

electrically-charged membranes in terms of their high permeability and selectivity. The

overall objective of this thesis was to develop a more fundamental understanding of the

performance characteristics of zwitterionic membranes by comparing the properties of a

series of charged, neutral, and zwitterionic membranes with similar pore size /

permeability.

The surface-modified membranes were generated by covalent attachment of small

ligands to a base cellulose membrane. The membranes were characterized using

streaming potential measurements, XPS, protein fouling, and protein transmission. The

latter was examined using lysozyme and α-lactalbumin charge ladders, which consist of a

series of chemical derivatives of the base protein differing by single charge groups that can

be analyzed by capillary electrophoresis. The sieving coefficients were analyzed using

available hydrodynamic models based on the partitioning of a charged sphere in a

charged cylindrical pore.

The zwitterionic membranes showed minimal protein adsorption and a very low

degree of protein fouling over a wide range of conditions. The fouling behavior of the

charged membranes was strongly influenced by electrostatic interactions, although the

zwitterionic membrane had a lower degree of fouling than the charged membranes even

iv

when the protein and membrane had like polarity. The low protein fouling characteristics,

coupled with their high selectivity, makes these zwitterionic membranes promising

candidates for high performance ultrafiltration processes.

v

Table of Contents

LIST OF FIGURES ....................................................................................................... ix

LIST OF TABLES ....................................................................................................... xii

Chapter 1 .........................................................................................................................1

1.1 Background .......................................................................................................1

1.2 Membrane technology for protein purification ...................................................2

1.3 Thesis Program ..................................................................................................6

Chapter 2 .........................................................................................................................8

2.1 Introduction .......................................................................................................8

2.2 Bulk Mass Transport .........................................................................................9

2.2.1 Concentration Polarization - Stagnant Film Model......................................9

2.2.2 Bulk Mass Transfer Coefficient ................................................................ 12

2.3 Membrane Transport ....................................................................................... 13

2.3.1 Solvent Transport - Membrane Hydraulic Permeability ............................ 13

2.3.2 Solute Transport - Thermodynamic Contributions .................................... 14

2.4 Protein Net Charge Analysis ............................................................................ 19

2.4.1 Protein Charge Calculations from Amino Acid Composition .................... 19

2.4.2 Protein Charge from Capillary Electrophoresis-Electrophoretic Mobility .. 21

Chapter 3 ....................................................................................................................... 23

vi

3.1 Introduction ..................................................................................................... 23

3.2 Membranes ...................................................................................................... 23

3.2.1 Membrane Materials ................................................................................. 23

3.2.2 Membrane Modification ........................................................................... 25

3.2.3 Streaming potential measurements ............................................................ 28

3.3 Solution Preparation ........................................................................................ 31

3.3.1 Buffer solutions ........................................................................................ 31

3.3.2 Protein Solutions ...................................................................................... 32

3.3.3 Dextran solutions...................................................................................... 34

3.4 Ultrafiltration .................................................................................................. 35

3.4.1 Apparatus ................................................................................................. 35

3.4.2 Membrane Hydraulic permeability ........................................................... 36

3.4.3 Protein Sieving ......................................................................................... 37

3.4.4 Diafiltration .............................................................................................. 38

3.4.5 Protein Fouling ......................................................................................... 38

3.5 Size Exclusion Chromatography (SEC) ........................................................... 39

3.6 Capillary electrophoresis ................................................................................. 40

3.7 X-ray Photoelectron Spectroscopy (XPS) ........................................................ 41

Chapter 4 ....................................................................................................................... 42

4.1 Introduction ..................................................................................................... 42

vii

4.2 Materials and Methods .................................................................................... 42

4.3 Results and Discussions ................................................................................... 43

4.3.1 Membrane Modification ........................................................................... 43

4.3.2 Membrane Surface Charge Characteristics................................................ 46

4.3.3 Dextran Ultrafiltration .............................................................................. 50

4.3.4 Static adsorption ....................................................................................... 51

4.3.5 IgG ultrafiltration ..................................................................................... 54

4.3.6 Protein ultrafiltration ................................................................................ 62

4.4 Conclusions ..................................................................................................... 64

Chapter 5 ....................................................................................................................... 66

5.1 Introduction ..................................................................................................... 66

5.2 Materials and Methods .................................................................................... 67

5.3 Results and Discussions ................................................................................... 67

5.3.1 Membrane Characterization ...................................................................... 67

5.3.2 Charge ladder Characterization ................................................................. 68

5.3.3 Ultrafiltration Experiments ....................................................................... 73

5.4 Conclusions ..................................................................................................... 77

6.1 Introduction ..................................................................................................... 80

6.2 Protein Transport through Surface Modified Membranes ................................. 81

6.3 Protein Fouling of Surface Modified Membranes............................................. 82

viii

6.4 Recommendations ........................................................................................... 83

References ..................................................................................................................... 85

Appendix ....................................................................................................................... 93

ix

LIST OF FIGURES

Figure 2.1 Schematic of concentration polarization during protein ultrafiltration. ...... 10

Figure 3.1 Molecular structure of cellulose. .............................................................. 24

Figure 3.2 Scanning electron micrograph showing the cross section of the composite

regenerated cellulose membrane. Taken from Burns (2000) with

permission. .............................................................................................. 25

Figure 3.3 Schematic of the reaction chemistry used to generate the surface modified

cellulose membranes (second reaction shown with the zwitterionic ligand).

................................................................................................................ 27

Figure 3.4 Molecular structure of the chemically-modified membranes where R is the

glucose monomer in the base cellulose. .................................................... 28

Figure 3.5 Schematic of the streaming potential apparatus used to determine the

effective membrane surface charge. ......................................................... 30

Figure 3.6 Schematic representation of the acylation reaction using acetic anhydride

(reproduced with permission from Ebersold and Zydney, 2004). .............. 34

Figure 3.7 Schematic of experimental set-up for constant pressure ultrafiltration

experiments.............................................................................................. 36

Figure 4.1 XPS spectra showing the nitrogen peak for all 100 kDa modified

membranes. .............................................................................................. 44

x

Figure 4.2 XPS spectra showing the sulfur peak for all 100 kDa modified membranes.

................................................................................................................ 44

Figure 4.3 Streaming potential data for 100 kDa Ultracel™ zwitterionic membrane in

10 mM buffered KCl solutions at several pH. ........................................... 48

Figure 4.4 Correlation between the apparent zeta potential and the calculated charge

based on the pKa values of the lysine ligand. ............................................ 49

Figure 4.5 Dextran sieving coefficients in 150 mM ionic strength at pH 7 using

different surface-modified membranes. .................................................... 51

Figure 4.6 Amount of protein adsorption from 5 g/L IgG at pH 9 on different modified

membranes. .............................................................................................. 52

Figure 4.7 Filtrate flux as a function of time during a typical fouling experiment for

the zwitterionic membrane using a 5 g/L IgG solution in 10 mM buffered

KCl at pH 7 and a constant pressure of 69 kPa (10 psi). ........................... 55

Figure 4.8 Filtrate flux (top panel) and filtrate concentration (bottom panel) for

ultrafiltration of a 5 g/L IgG solution in 10 mM buffered KCl at pH 5 at a

constant pressure of 69 kPa through the zwitterionic, positive, and negative

(sulfonic acid) membranes. ...................................................................... 59

Figure 5.1 Capillary electropherograms for α-lactalbumin (top panel) and lysozyme

(bottom panel) charge ladders in 10 mM tris/glycine buffer at pH 8.3. ..... 70

xi

Figure 5.2 Net charge for the first seven peaks in the α-lactalbumin and lysozyme

charge ladders evaluated both from the electrophoretic mobility data using

10 mM tris/glycine buffer at pH 8.3 (open symbols) and from the amino

acid composition (filled symbols)............................................................. 73

Figure 5.3 Observed sieving coefficients for ultrafiltration of lysozyme charge ladder

at pH 7 through 100 kDa modified Ultracel™ membrane as a function of

net protein charge. .................................................................................... 75

Figure 5.4 Observed sieving coefficients as a function of charge interaction parameter

(product of the dimensionless surface charge densities for the protein and

the membrane) for different membranes. Filled symbols represent the

experimental data with the solid curves representing the theoretical model.

................................................................................................................ 77

xii

LIST OF TABLES

Table 2.1 Expansion coefficients for Kt and Ks functions in Equation 2.9 and 2.10.16

Table 3.1 Physicochemical properties of proteins. ..................................................... 33

Table 4.1 Atomic composition (percent) and calculated degree of modification for

different 100 kDa modified membranes determined from XPS data. .......... 46

Table 4.2 The experimental values for the apparent zeta potential of surface-modified

100 kDa Ultracel™ membranes at pH 7 in 10 mM buffered KCl. .............. 50

Table 4.3 Effect of protein adsorption on the permeability of the surface-modified 100

kDa Ultracel™ membranes........................................................................ 54

Table 4.4 Buffer flux (after protein adsorption) and initial filtrate flux with a 5 g/L IgG

solution for the different membranes at pH 5, 7, and 9. .............................. 57

Table 4.5 Flux recovery of different modified 100 kDa membranes after 1 hr

ultrafiltration of an IgG solution at pH 5, 7, and 9. ..................................... 61

Table 4.6 Initial buffer flux and flux recovery ratio for the zwitterionic and positively-

charged membranes after ultrafiltration of a 5 g/L IgG solution at pH 5 with

and without a pre-adsorption step. ............................................................. 62

Table 4.7 Flux recovery of zwitterionic and positively-charged 100 kDa membranes

after 1 hr ultrafiltration of lysozyme and α-lactalbumin at pH 7 and BSA at

pH 4.7. ...................................................................................................... 63

xiii

Table 5.1 Apparent zeta potential of surface-modified 100 kDa Ultracel™ membranes

at pH 7 in 10 mM buffered KCl. ................................................................ 68

1

Chapter 1

Introduction

1.1 Background

The production of high value recombinant proteins requires robust, cost-effective,

and high-resolution purification methods that can provide high yield and purification of

the desired product. Although there can be considerable variability in the economics for

different therapeutic proteins, several studies have reported that up to 80% of the total

manufacturing cost is associated with the downstream purification process (Clark and

Blanch, 1997). This is most pronounced in the production of high-dose therapeutic

proteins, e.g., monoclonal antibodies used in the treatment of various forms of cancer and

a number of immunologic disorders (van Reis and Zydney, 2007; Zydney, 2009). Current

annual demand for some antibody products is on the order of 1000 kg, creating a number

of new separation challenges for the downstream purification process (Thömmes and

Etzel, 2007; Zydney, 2009). There is tremendous interest in the development of improved

separation technologies to meet the needs of the evolving biotechnology industry.

2

1.2 Membrane technology for protein purification

Membranes are very attractive for the separation and purification of high value

proteins since they operate under mild conditions that will not degrade or damage the

biological products. In addition, membranes have robust performance, high throughput,

easy scale-up, low energy consumption, low space requirements, and environmentally

friendly operating characteristics (no need for toxic chemicals) (Guo et al., 2012; Li et al.,

2012a). Ultrafiltration (UF) membranes with pore size from around 1 to 15 nm are widely

used for protein concentration and buffer exchange; UF is the method of choice for final

formulation of nearly all recombinant protein products (van Reis and Zydney, 2007).

UF was originally viewed as a purely size-based separation process with the

proteins retained by the membrane due to steric exclusion from the small pores.

However, as early as 1975, Chang et al. (Chang et al., 1975) reported that the

transmission of a synthetic polyanion (dextran sulfate) through the renal glomerular

capillary was about 20-fold smaller than that of a neutral dextran of similar size and

structure. Malone and Anderson (Malone and Anderson, 1978) attributed the observed

reduction in the hindered diffusion coefficient of latex particles through track-etched

mica membranes at low salt concentration to strong electrostatic interactions at low ionic

strength. Subsequent studies have extended these observations to pressure-driven

filtration of charged proteins through a range of commercial ultrafiltration membranes

with different surface charge characteristics and under different solution conditions. For

example, Mehta and Zydney (Mehta and Zydney, 2006) showed that the transmission of

a positively-charged protein through a series of positively charged membranes was a

3

strong function of membrane charge. Pujar and Zydney (Pujar and Zydney, 1994)

reported that the transmission of negatively charged bovine serum albumin through a

negatively charged membrane decreased by more than two orders of magnitude at low

ionic strength due to the absence of significant electrostatic shielding under these

conditions. These results clearly demonstrate that protein transport through

semipermeable ultrafiltration membranes is strongly affected by electrostatic interactions

between the charged membrane and the charged protein.

Electrostatic effects have also been exploited to achieve high resolution protein

separations, with the charged membranes providing high retention of like-charged

proteins while allowing relatively uncharged solutes to pass into the permeate (van Reis

and Zydney, 2007). Examples include the separation of bovine serum albumin (BSA)

from hemoglobin (van Eijndhoven et al., 1995), the separation of BSA and

immunoglobulin G (Saksena and Zydney, 1994), the purification of an antigen binding

fragment from BSA (Van Reis et al., 1999), the purification of a monoclonal antibody

from Chinese Hamster Ovary host cell proteins (van Reis and Zydney, 2007), and the

purification of an antibody fragment from E. Coli host cell proteins (Lebreton et al., 2008).

Electrostatic interactions can also be used to develop high performance ultrafiltration

membranes with significantly greater protein retention (or selectivity) for a given value of

the permeability than conventional UF membranes (Mehta and Zydney, 2006; Zydney,

2009).

One of the challenges in many UF processes is membrane fouling due to specific

interactions between the proteins and membrane surface. Fouling causes a decline in the

filtrate flux, increasing the required feed pressure and requiring additional membrane

4

cleaning (Guo et al., 2012). A variety of strategies are used to control / minimize fouling

including modification of the membrane surface to reduce protein – membrane interactions

(Pasmore et al., 2001; Vrijenhoek et al., 2001). Several recent studies have focused on the

development of non-porous zwitterionic surfaces, analogous to the zwitterionic character

(presence of both positive and negative charge groups in close proximity) of phospholipid

cell membranes (Chen et al., 2005; Ishihara and Takai, 2009; Lewis, 2000; Vermette and

Meagher, 2003). The low fouling behavior of these zwitterionic materials is attributed to

their ability to bind a significant number of water molecules through both electrostatic

interactions and hydrogen bonding (Chen et al., 2005; Harder et al., 1998; Ishihara et al.,

1998; Li et al., 2012b; Nagumo et al., 2012; Sun et al., 2006). This hydration layer

effectively excludes the protein from the surface thus preventing adsorption (He et al., 2008;

Ishihara et al., 1998).

The low fouling characteristic of zwitterionic surfaces has also been exploited in the

development of porous membranes. For example, Jiang and co-workers showed that

membranes made from a zwitterionic sulfobetaine copolymer with either polyethersulfone

(Wang et al., 2006) or acrylonitrile (Sun et al., 2006) were highly hydrophilic and more

resistant to protein fouling than the base membranes during ultrafiltration of bovine serum

albumin (BSA). Li et al. (Li et al., 2012b) obtained similar results by grafting the

zwitterionic monomer sulfobetaine methacrylate onto the surface of poly(vinylidene

fluoride) (PVDF) membranes. An et al. (An et al., 2013) used zwitterionic amine monomers

to make thin-film composite nanofiltration membranes. The relative flux recovery ratio after

BSA ultrafiltration increased from 0.83 to 0.93 by increasing the zwitterionic monomer

concentration from 0 to 3.2 mol%. Ji et al. (Ji et al., 2012) introduced a novel zwitterionic

5

terpolymer to nanofiltration membranes, significant increasing the water flux recovery after

BSA ultrafiltration. Zwitterionic grafted PVDF hollow fiber membranes show high

hydrophilicity and strong resistance to both protein and oil fouling (Li et al., 2012c).

Zwitterionic grafted PES hollow fiber membranes have shown improved antifouling

behavior in comparison with the un-modified membrane for both BSA and lysozyme (Razi

et al., 2012).

In addition to the use of long-chain zwitterionic polymers, several studies have

demonstrated the potential of using small zwitterionic ligands, e.g., peptides and amino

acids, for membrane surface modification (Chelmowski et al., 2008; Shi et al., 2011). For

example, Shi et al. (Shi et al., 2011) grafted the amino acids lysine, glycine, and serine onto

the surface of a hydrolyzed polyacrylonitrile membrane containing a high concentration of

carboxylic acid groups. All of the modified membranes had similar hydrophilicity, as

determined from water contact angle measurements, but the lysine-modified membranes

showed the least protein fouling under both static and dynamic (filtration) conditions.

However, it is difficult to interpret these results since all of the membranes had a significant

negative charge due to the presence of the carboxylic acids in the hydrolyzed

polyacrylonitrile membrane.

Although these recent studies have clearly demonstrated the potential of zwitterionic

ultrafiltration membranes, there are still considerable uncertainties regarding the factors

controlling the performance of these membranes and how they compare to more

conventional charged and neutral membranes. Most of the early studies in this area were

performed using zwitterionic polymers, in which case the modified membranes tended to

have very different permeability, pore size, and even surface morphology than the

6

unmodified membranes. The work by Shi et al. (Shi et al., 2011) used small ligands to

eliminate many of these effects, but the base membrane had a significant negative charge

which persisted in the modified membranes. In addition, none of these studies examined

protein transmission through the zwitterionic membranes, making it impossible to

determine whether these low fouling membranes had the combination of selectivity and

permeability that is needed for high performance ultrafiltration processes.

1.3 Thesis Program

The overall objective of this thesis was to develop a more fundamental

understanding of the performance characteristics of zwitterionic membranes by comparing

the properties of these membranes with a series of charged and neutral membranes with

similar pore size / permeability. This included: (1) studying the fouling characteristics of

these membranes during both static protein adsorption and actual ultrafiltration over a

range of solution pH, and (2) investigating the effects of electrostatic interactions on

protein transport through these membranes using protein charge ladders, which consist of

a series of chemical derivatives of a given protein differing only by single charge groups.

The general theoretical background used to analyze the ultrafiltration results is

presented in Chapter 2. This includes a brief review of available theoretical analyses for

solute and solvent transport through membranes with relatively small pores, with a

particular emphasis on the effects of electrostatic interactions on solute partitioning into

7

charged membrane pores. The last part reviews the calculation of the protein net charge

from both the electrophoretic mobility and the known amino acid sequence.

Chapter 3 describes the experimental set-up, materials, and methods used in the

experimental studies described in this thesis. Specific details on some of the

experimental procedures are provided in the appropriate Chapters.

Chapter 4 presents membrane fouling studies under both static exposure and

dynamic ultrafiltration using human serum IgG as a model protein. Limited data are also

provided with BSA, α-lactalbumin, and lysozyme.

Chapter 5 examines the transport of protein charge ladders through zwitterionic,

positively-charged, acidic, and hydrophilic membranes generated by covalent

modification of a base cellulose membrane using specific small ligands. Ultrafiltration

data were analyzed using available theoretical models describing the partitioning of a

charged sphere in a charged cylindrical pore.

Chapter 6 summarizes the major contributions of this thesis and makes several

recommendations for future studies on the development of high performance zwitterionic

ultrafiltration membranes for protein purification.

8

Chapter 2

Theoretical Background

2.1 Introduction

This Chapter provides a brief review of the theoretical models developed to

describe the basic mass transport and separation phenomena governing the behavior of

ultrafiltration systems with an emphasis on electrostatic interactions. Previous reviews of

this theoretical analysis have been presented by Zeman and Zydney (Zeman and Zydney,

1996) and in several dissertations published under the direction of Professor Andrew

Zydney (Burns, 2000; Mehta, 2006; Molek, 2008). The discussion below draws

extensively from these prior reviews.

The overall rate of protein transport through a semipermeable ultrafiltration

membrane is determined by the rate of protein transport from the bulk solution to the

membrane and through the membrane pores. Transport in the bulk solution is governed

primarily by the system hydrodynamics while transport through the membrane pores has

contributions from both thermodynamics (including electrostatic interactions) and

hydrodynamics.

9

2.2 Bulk Mass Transport

2.2.1 Concentration Polarization - Stagnant Film Model

The pressure-driven flow through a semipermeable membrane causes an

accumulation of the completely or partially retained solute at the upstream surface of the

membrane. This phenomenon is called concentration polarization and causes the protein

(solute) concentration to vary from the value in the bulk solution ( ) to a much greater

value at the membrane surface/wall ( over the distance of the concentration

polarization boundary layer thickness, . Figure 2.1 shows a schematic of the

concentration polarization phenomenon, including the expected concentration profile.

The high concentration at the upstream surface of the membrane can significantly affect

the overall system behavior. It increases the driving force for protein transport into and

through the membrane, while reducing the effective pressure driving force for solvent

transport due to the osmotic pressure associated with the retained protein. The high

protein concentration at the membrane surface can also contribute to an increase in

membrane fouling.

10

Figure 2.1 Schematic of concentration polarization during protein ultrafiltration.

Concentration polarization in membrane systems is most commonly analyzed

using the stagnant film model. This model approximates the concentration profile

upstream of the membrane by treating the problem as a one-dimensional (stagnant)

boundary layer, neglecting the complexities associated with the detailed fluid flow

characteristics in the particular module as well as the coupling between mass and

momentum transport. In the classical model, protein-protein interactions are neglected

and the solute diffusivity and viscosity are both assumed to be independent of the solute

concentration and constant throughout the boundary layer. At steady state, the net solute

flux towards the membrane is set equal to the solute flux through the membrane and into

the filtrate solution:

[2.1]

𝐶𝑏

𝐶𝑤

𝐶𝑓

11

where is the filtrate flux through the membrane (equal to the volumetric filtrate flow

rate per unit membrane area), is the concentration of solute in the filtrate solution, is

the local solute concentration at a position above the membrane surface, and is the

free solution diffusion coefficient of the solute. Equation 2.1 is integrated across the

concentration boundary layer (from at to at ) giving the

following expression for the filtrate flux:

(

) [2.2]

A more detailed analysis of this model is provided by Zydney (Zydney, 1997).

Equation 2.2 can also be used to evaluate the effect of the filtrate flux on the

observed sieving coefficient (

. The observed sieving coefficient is typically

evaluated in terms of the actual sieving coefficient, which is defined as ratio of the solute

concentration in the filtrate to that at the membrane wall ( as:

(

)

(

) [2.3]

At low filtrate flux (

), the observed sieving coefficient is equal to the actual sieving

coefficient since concentration polarization is minimal. increases with increasing filtrate

flux, approaching a value of one as .

12

2.2.2 Bulk Mass Transfer Coefficient

The solute mass transfer coefficient, , is defined as the ratio of the solution

diffusion coefficient ( ) to the boundary layer thickness ( ) in Equations 2.2 and 2.3.

This coefficient is a function of the solute diffusivity and the hydrodynamics of the

particular membrane device. Although it is possible to evaluate the mass transfer

coefficient theoretically for relatively simple module configurations, the analysis of mass

transfer in the stirred cells used in this research is difficult due to the complex flow

profiles in this system. Instead, the mass transfer coefficient is typically evaluated using a

semi-empirical equation for mass transfer in a stirred cell for laminar flow (

) that was developed by Smith et al. (Smith et al., 1968) based on the rate of

benzoic acid dissolution into a stirred solution as:

[2.4]

where

is the Sherwood number,

is the Reynolds number,

is

the Schmidt number, is the radius of the stirred cell, is the stirring speed, and is the

kinematic viscosity. Opong and Zydney (Opong and Zydney, 1991) evaluated as 0.23

for the 25 mm diameter Amicon stirred cell used in this thesis based on data for the filtrate

flux as a function of the transmembrane pressure at several bulk protein concentrations and

stirring speeds. The protein diffusion coefficient ( in m2/s) can be evaluated as (Young

et al., 1980):

[2.5]

13

where is the solution viscosity (in Pa s), is the absolute temperature (in K), is the

Boltzmann’s constant , and is the protein molecular weight in

g/mol. Equation 2.5 is valid only at infinite dilution because it neglects the effects of

protein-protein interactions.

2.3 Membrane Transport

The rate of solute and solvent transport through porous membranes is typically

described using hydrodynamic theories in which the membranes are modeled as an array

of well-defined cylindrical pores, while the solutes are considered to behave as uniform

rigid spheres (Anderson and Quinn, 1974; Deen, 1987). The advantage of the

hydrodynamic models is that the key transport parameters can be calculated directly in

terms of the physical properties of the solute and the pores. Hydrodynamic theories can

be easily extended to incorporate the effects of a pore size distribution (by numerical

integration over the distribution) as well the effects of electrostatic interactions

(Mochizuki and Zydney, 1993; Saksena and Zydney, 1995).

2.3.1 Solvent Transport - Membrane Hydraulic Permeability

The rate of solvent transport through a membrane is generally described in terms

of the hydraulic permeability ( ):

14

[2.6]

where is the transmembrane pressure.

The rate of solvent transport is also dependent on the membrane surface charge

and solution ionic strength due to electrokinetic effects. The presence of a net surface

charge on the pore wall causes an accumulation of counter-ions in the electrical double

layer adjacent to the pore wall. The pressure-driven convective fluid flow through the

charged pore will generate an unequal flux of the co-ions and counter-ions, leading to the

development of an induced (streaming) potential. At steady state, the streaming potential

generates a back conductive ion transport that exactly balances the convective ion flux,

resulting in a situation in which there is no net current flow through the pore. The

induced streaming potential reduces the magnitude of the solvent flux due to the net force

on the fluid exerted by the action of the electric field on the ions (often referred to as

counter-electroosmosis). A detailed review of solvent transport through electrically-charged

membranes is provided elsewhere (Burns, 2000; Pujar and Zydney, 1994).

2.3.2 Solute Transport - Thermodynamic Contributions

The rate of protein transport through small pore ultrafiltration membranes is

typically analyzed in terms of both thermodynamic and hydrodynamic interactions, with

the actual protein sieving coefficient ( ) expressed as:

[2.7]

15

where is the thermodynamic equilibrium partition coefficient between the bulk solution

and the membrane pore and is the hindrance factor for convection, which accounts for the

additional hydrodynamic drag on the solute molecule due to the presence of the pore wall.

Equation 2.7 assumes that protein transport is dominated by convection, which is a

reasonable approximation during protein ultrafiltration due to the relatively high Peclet

numbers in these systems.

The hindrance factor for convection, , can be evaluated in terms of an integral

over the radial coordinate in the pore. For a solute located at the pore axis (i.e., at the pore

centerline), the integral becomes (Deen, 1987):

[2.8]

where is the ratio of the solute to pore radius. Expressions for Ks and Kt can be

developed using matched asymptotic expansions (Bungay and Brenner, 1973) giving:

[ ∑ ] ∑

[2.9]

[ ∑ ] ∑

[2.10]

with the expansion coefficients provided in Table 2.1.

16

Table 2.1 Expansion coefficients for Kt and Ks functions in Equation 2.9 and 2.10.

Subscript, n

1 -73/60 7/60

2 77,293/50,400 -2,227/50,400

3 -22.5083 4.0180

4 -5.6117 -3.9788

5 -0.3363 -1.9215

6 -1.216 4.392

7 1.647 5.006

The equilibrium partition coefficient is defined as the ratio of the average protein

concentration in the pore to that in the bulk external solution immediately adjacent to the

membrane:

∫ [

]

[2.11]

where is the pore radius, is the radial coordinate within the cylindrical pore, is the

Boltzmann constant, is the absolute temperate, and is the total interaction

potential. Smith and Deen (Smith and Deen, 1980) developed the first rigorous analytical

expressions for the electrostatic potential for a spherical solute in a cylindrical pore by

solving the linearized Poisson-Boltzmann equation using matched asymptotic expansions

17

in cylindrical and spherical coordinates. The results for interactions at constant surface

charge density are conveniently expressed as:

[2.12]

The coefficients , , , and are all positive functions of the solution ionic

strength, solute radius, and pore radius:

[2.13]

[2.14]

[2.15]

[2.16]

[2.17]

∫ [(

) ]

[

]

[2.18]

where and are modified Bessel functions, is the dimensionless pore radius,

and is the inverse Debye length:

[ ∑

]

[2.19]

where and are the valence and concentration of each ion.

18

The dimensionless surface charge densities of the solute, , and pore, , are

defined as:

[2.20]

[2.21]

with the permittivity of free space , the dielectric constant of

the solution, the Faraday constant , and the ideal gas constant

. The dimensional surface charge densities of the solute and pore (

and ) can be evaluated in terms of the net electronic charge on the solute (Z) and the

apparent zeta potential of the membrane pore ( ), respectively, as:

[2.22]

(

) [2.23]

where is the electron charge and is the bulk electrolyte

concentration.

Equations 2.13 to 2.18 are valid for a solute located at the pore axis.

Corresponding results are available for arbitrary radial positions (Smith and Deen, 1983)

as well as for interactions at constant surface potential instead of constant surface charge

density (Smith and Deen, 1980, 1983). It is also possible to extend this analysis to

account for the effects of charge regulation using a linearized form of the charge

regulation boundary condition; the resulting equations account for the change in surface

19

charge / potential of the protein and pore wall associated with the alteration in the local

electrical potential field (and ion concentrations) when the protein enters the pore (Pujar

and Zydney, 1997).

2.4 Protein Net Charge Analysis

2.4.1 Protein Charge Calculations from Amino Acid Composition

The net electrical charge on a protein is determined by its amino acid sequence

due to the dissociation of the various ionizable residues on the surface of the protein

along with the adsorption (or binding) of specific ions from the bulk electrolyte. The

number of dissociated acidic and basic amino acid residues can be calculated from the

intrinsic dissociation constant for each amino acid. For example, the dissociation

equilibrium for an α-carboxylic acid can be expressed as:

[ ][ ]

[ ] [2.24]

Equation 2.24 can be rewritten in terms of the pH and the number of dissociated groups

( ) as:

[2.25]

where [ ], [

], and is the total number of titratable

species. The concentration at the protein surface (required in Equations 2.24 and 2.25)

is different from the bulk concentration due to electrostatic interactions between the

20

charged protein and the charged hydrogen ion. The local concentration at the membrane

surface can be related to the bulk concentration using a classical Boltzmann distribution:

[2.26]

where is the bulk hydrogen ion concentration, is the electron charge, and is the

electrostatic potential at the protein surface:

[2.27]

where is the net protein surface charge (evaluated as the number of electronic charges).

Equation 2.27 is developed assuming that the protein is a hard sphere with the electrical

charges distributed uniformly over the spherical surface (Overbeek and Wiersema, 1967).

The protein charge is equal to the difference between the maximum number of positive

charges (N-terminal, histidine, lysine, and arginine) and the sum of all the dissociated

groups:

[2.28]

Equations 2.26 to 2.28 can be solved iteratively to evaluate the net protein charge

as a function of the bulk pH and solution ionic strength (which determines the Debye

length). The development of these equations is discussed in more detail by Menon and

Zydney (Menon and Zydney, 2000). The number and values of the various amino

acids present in the model proteins used in this thesis are summarized in Appendix.

21

2.4.2 Protein Charge from Capillary Electrophoresis-Electrophoretic

Mobility

The electrophoretic mobility reflects the balance between the electrical forces

arising from the applied electric field and the hydrodynamic (friction) forces associated

with the viscosity of the suspending medium. A large number of theoretical models have

been developed for the electrophoretic mobility, with the differences primarily in the

approximations made in evaluating the electrical interactions. These approximations

include both the detailed structure of the equilibrium electrical potential (e.g., the use of

the low electrical potential, small Debye length, or flat plate approximations) and the

evaluation of the distortion of the equilibrium structure associated with the particle and

fluid motion during electrophoresis.

A simple analytical expression for the electrophoretic mobility in terms of the

electrical potential at the surface of a hard sphere for the limiting case when the electrical

double layer thickness is much smaller than the particle radius ( ) and the electrical

potential is relatively small (Debye-Huckel approximation) is given as (Overbeek and

Wiersema, 1967):

[2.29]

where is the electrostatic potential at the particle surface, is the electrical permittivity

of the solution, and is the solution viscosity.

22

If the electrical double layer is much larger than the particle radius ( ), the

electrophoretic mobility is given by the Debye-Huckel equation assuming that the

potential is low:

[2.30]

Henry (Henry, 1931) obtained a more complete solution for the electrophoretic mobility

that accounts for the distortion of the electric field lines by the presence of the particle,

with the resulting expression valid over the entire range of Debye lengths:

[2.31]

with

[

] ∫

[2.32]

where is Henry’s function, which accounts for finite double layer thickness, and is a

dummy variable over which the integration is performed. The zeta potential for a

uniformly charged hard sphere can be expressed in terms of the particle net charge using

Equation 2.27. The net electrical charge of the protein is thus given as:

[2.33]

More details on the evaluation of the electrophoretic mobility from capillary

electrophoresis experiments are provided elsewhere (Menon, 1999; Molek, 2008).

23

Chapter 3

Materials and Methods

3.1 Introduction

This chapter describes the materials, apparatus, and methods used for the

experimental studies performed in this thesis. Additional details on specific materials or

methods are provided in subsequent chapters as appropriate.

3.2 Membranes

3.2.1 Membrane Materials

Asymmetric membranes are used in almost all commercial applications of

ultrafiltration. These membranes are anisotropic and have a thin skin, which provides the

membrane its functionality, and a much thicker and more porous support that provides the

membranes its structural integrity. The small thickness of the skin allows much higher

fluxes to be obtained compared to symmetric membranes with comparable selectivity.

Although a variety of polymers can be used to make asymmetric ultrafiltration

membranes, regenerated cellulose is one of the most attractive membrane materials.

24

Cellulosic membranes have become a major component of the downstream purification

process for therapeutic proteins in the biotechnology industry. The free hydroxyl groups on

the glucose rings within the cellulose polymer renders the membrane highly hydrophilic,

significantly reducing protein binding and fouling during use. These hydroxyl groups are

also available for chemical modifications as discussed in the next section. Figure 3.1 shows

a schematic of the chemical structure of cellulose.

Figure 3.1 Molecular structure of cellulose.

Ultrafiltration experiments were performed using UltracelTM

composite

regenerated cellulose membranes with nominal molecular weight cut-off (MWCO) of

100 kDa provided by Millipore Corp. (Bedford, MA). UltracelTM

membranes with 10

kDa molecular weight cut-off were used for buffer exchange. The nominal molecular

weight cut-off refers to the molecular weight of a solute which has approximately 90%

rejection as determined by the manufacturer. A scanning electron micrograph (SEM) of

the cross section of the composite regenerated cellulose membrane is shown in Figure

3.2; the skin layer (approximately 0.5 - 1 μm thick) is not visible in the SEM. The two

layers seen in the SEM correspond to the porous cellulosic substructure (approximately

25

60 μm thick) and the porous polyethylene substrate on which the cellulose membrane is

cast.

Figure 3.2 Scanning electron micrograph showing the cross section of the composite

regenerated cellulose membrane. Taken from Burns (Burns, 2000) with

permission.

Membrane disks with 25 mm diameter were cut from large flat sheets using a

stainless-steel cutting device fabricated in our laboratory. All membranes were soaked in

90% (V/V) isopropanol for 45 min to remove any protective agents. The membranes

were then thoroughly rinsed with at least 100 L/m2 of deionized (DI) water.

3.2.2 Membrane Modification

Most of the approaches used for surface modification of cellulose membranes are

based on the activation and subsequent reaction of the free hydroxyl groups on the base

cellulose. Surface-modified ultrafiltration membranes were prepared from the Ultracel™

composite regenerated cellulose membranes by covalent attachment of different ligands

26

to epoxy-activated hydroxyl groups using the reaction chemistry developed by Liu et al.

(Liu et al., 2005) and subsequently modified by Mehta and Zydney (Mehta and Zydney,

2008) to avoid degradation of the cellulose (shown schematically in Figure 3.3). All

membranes were generated using the same linkage chemistry but with different ligands

having very similar chain length but with different end functionality. This made it

possible to study the effects of ligand chemistry on membrane surface characteristics

independent of the membrane pore size, providing an appropriate set of controls for

subsequent studies.

Membranes were first incubated in 0.1 M NaOH for 24 hr. The cellulose surface

was then activated by incubating the membrane in a 25 mL capped plastic jar for 2 hr at

45 oC in a mixture of epichlorohydrin (Alfa Aesar, A15823) and 0.1 M NaOH in a 1:2

ratio (by volume) (step 1 in Figure 3.3). The membranes were then carefully removed and

rinsed with DI water. The epoxide groups on the activated membranes were reacted with

the desired ligand (step 2 in Figure 3.3) by immersing in 20 mL of a 1 M solution of that

ligand: L-Lysine (Sigma, L5501) for the zwitterionic surface, hexamethylenediamine

(Sigma, H11696) for the positively-charged surface, 6-aminocaproic acid (Sigma,

A2504) for the acidic surface, 6-amino-1-hexanol (Sigma, A56353) for the hydrophilic

surface, and hexylamine (Sigma, 219703) for the hydrophobic surface. The molecular

structures of the different membranes are shown schematically in Figure 3.4 where R is

the glucose ring of the cellulose membrane. All reactions were conducted at 45°C for 12

hr to insure complete reaction with the epichlorohydrin groups. The membranes were

then removed from the reaction solution and thoroughly rinsed with DI water. Data were

also obtained with a negatively charged version of the Ultracel™ membrane generated by

27

direct chemical attachment of a ligand containing sulfonic acid groups to the free

hydroxyl of the cellulose (without activation with epichlorohydrin) as described

elsewhere (Van Reis, 2006).

Figure 3.3 Schematic of the reaction chemistry used to generate the surface modified

cellulose membranes (second reaction shown with the zwitterionic ligand).

28

Figure 3.4 Molecular structure of the chemically-modified membranes where R is the

glucose monomer in the base cellulose.

3.2.3 Streaming potential measurements

The effective surface charge of the different ultrafiltration membranes was

evaluated from streaming potential measurements using the approach described by Burns

and Zydney (Burns and Zydney, 2000). The membrane was placed between two

Plexiglass chambers (Figure 3.5), each filled with 10 mM buffered KCl solution at the

desired pH, taking care to remove all air bubbles. Ag/AgCl electrodes were then screwed

tightly into the ends of the chambers to ensure reproducible placement (approximately

29

0.1 – 0.2 cm away) relative to the membrane surface (with O-rings in place to provide

good seals).

The electrodes were prepared as follows. 1 mm diameter silver wires (Sigma

Chemical Co., St. Louis, MO) were lightly sanded and placed in a concentrated nitric

acid solution for approximately 10 s. Each wire was then washed with DI water and

placed in a 1 M KCl solution. A DC power supply was then connected to each silver

electrode and a steel wire, the current was set at 20 mA, and a uniform Ag/AgCl layer

was deposited on the wire surface for 20 min. Electrodes were stored in 0.5 M KCl

solution between experiments.

To measure the streaming potential of the membranes, an air-pressurized feed

reservoir containing an appropriate buffer solution was attached to the feed chamber. The

electrodes were connected to a Keithley 2000 Multimeter to evaluate the transmembrane

voltage as a function of the transmembrane pressure ( ), which was set by

pressurizing the feed chamber. The system was allowed to stabilize for approximately 15

to 30 min at each pressure before evaluating the voltage. The system pressure was

gradually increased from 14 to 34 kPa (2 to 5 psi), with data obtained at four or more

discrete pressures. The apparent zeta potential was evaluated from the slope of the

voltage (streaming potential) as a function of pressure using the Helmholtz-

Smoluchowski equation (Hunter, 1981):

[3.1]

where is the solution conductivity, is the permittivity of free space, is the dielectric

constant of the solution, and is the solution viscosity. Note that Equation 3.1 is only valid

30

under conditions where the double layer thickness is very small compared to the pore radius.

Thus, should be considered an apparent or effective zeta potential, which is directly

related to the membrane surface charge but may not be equal to the actual value. A more

detailed discussion on the evaluation of the membrane charge from the streaming potential

is provided by Burns (Burns, 2000).

Figure 3.5 Schematic of the streaming potential apparatus used to determine the

effective membrane surface charge.

31

3.3 Solution Preparation

3.3.1 Buffer solutions

Different buffer solutions were prepared by dissolving preweighed amounts of the

appropriate salts in deionized distilled water obtained from a NANOpure® Diamond

water purification system (Barnstead Thermolyne Corporation, Dubuque, IA) with a

conductivity less than 56 nS/cm. KCl powder (BDH Chemicals, BDH0258) and either 1

mM acetate (Sigma, S7670), Bis-Tris (MPBiomedicals, 101038), or borate (Sigma,

S9640) were used for pH 5, 7, and 9, respectively. The running buffer for the capillary

electrophoresis was prepared by adding 192 mM glycine (Sigma, G7403) and 25 mM

Trizma® base (Sigma, T1503) to a 10 mM KCl solution at pH 8.3. Higher ionic strength

solutions were prepared by dissolving appropriate amounts of KCl in tris/glycine

solution. All salts were analytical reagent grade. The solution pH was measured using a

Model 402 Thermo Orion pH meter (Beverly, MA) and adjusted using 0.1 M sodium

hydroxide (NaOH) or hydrochloric acid (HCl) as required. A 105 A plus conductivity

meter (Thermo Orion, Beverly, MA) was used to measure the solution conductivity. All

buffer solutions were filtered through a 0.2 μm pore size Supor® 200 membrane (Pall

Corporation, Ann Arbor, MI) prior to use to remove any particles or un-dissolved salt.

The ionic strength of the buffer solution was evaluated as:

[3.2]

where and are the net charge and concentration of each ion, respectively.

32

3.3.2 Protein Solutions

Ultrafiltration experiments were performed with lysozyme and α-lactalbumin

charge ladders (described below). Fouling experiments were done using human serum

gamma globulin (IgG), bovine serum albumin (BSA), lysozyme, and α-lactalbumin as

model proteins. Table 3.1 summarizes the key physical properties and catalog numbers.

The protein isoelectric point and equivalent radius were taken from literature data. More

detailed information on the amino acid composition is provided in the Appendix.

Concentrated human serum IgG (obtained from SeraCare) was diluted with appropriate

buffer at the desired pH and ionic strength. Other protein solutions were prepared by

slowly dissolving pre-weighed amounts of protein powder in the desired buffer. The pH

of the protein solution was adjusted by adding appropriate amounts of 0.1 M acid or base

as required (e.g., HCl and KOH for buffered KCl solution). All protein solutions were

filtered through 0.22 μm Acrodisc® syringe filters (Pall Corporation) immediately prior

to use in the ultrafiltration experiments. Protein solutions were stored at 4°C and used

within 12 hr of preparation to minimize the likelihood of protein aggregation or

denaturation.

Protein concentrations were determined spectrophotometrically using a

SPECTRAmax Plus 384 UV-vis spectrophotometer (MD Corporation, Sunnyvale, CA)

with the absorbance evaluated at 280 nm for IgG and α-lactalbumin and 260 nm for

lysozyme. Actual concentrations were determined by comparison of the absorbance with

that of known protein standards.

33

Table 3.1 Physicochemical properties of proteins.

Protein

Molecular

weight,

(kDa)

Isoelectric

point

Equivalent

radius*, (nm)

Catalog

number

Lysozyme

(Rohani and Zydney, 2009)

14.3 11 1.6 Sigma

L6876

α-lactalbumin

(Molek and Zydney, 2007;

Rohani and Zydney, 2012)

14.2 4.6 1.6 Sigma

L5385

IgG

(Andersen et al., 2000;

Saksena and Zydney, 1994)

155 ≈7 5.5 Sera

Care

HS-475

BSA

(Menon and Zydney, 1998;

Razi et al., 2012)

67 4.7 3.45 Sigma

A7906

*Radius of sphere of equivalent volume

Protein charge ladders were synthesized by reaction of the lysine ɛ-amino groups

with acetic anhydride (Sigma, 242845) following the procedure described by Chung et al.

(Chung et al., 2009) as shown in Figure 3.6. The acetylated amide eliminates a potentially

protonable group by chemically blocking the lysine amino group. Thus, the resulting

charge ladder consists of a series of proteins with essentially the same size but each

differing by one or more charge groups. The lysozyme and α-lactalbumin charge ladders

were prepared by adding 1 M NaOH to a 10 g/L protein solution to bring the pH to 12.

Approximately 4 equivalents of 0.1 M acetic anhydride (per mole of protein) in 1,4-

dioxane (Sigma, 360481) were added to the solution, with the pH kept constant

throughout the 5 min reaction by addition of 0.1 N NaOH as needed. The pH was rapidly

34

lowered to 7 by addition of 1 M HCl to quench the reaction. The resulting protein

solution was diafiltered with at least four diavolumes of chilled DI water to remove the

dioxane, unreacted acetic anhydride, and any reaction by-products.

Figure 3.6 Schematic representation of the acylation reaction using acetic anhydride

(reproduced with permission from Ebersold and Zydney, (Ebersold and

Zydney, 2004)).

3.3.3 Dextran solutions

Dextrans are branched polymers made of glucose, joined by α-1,6 linkages in the

main chain and a small number of branches attached to the main chain by α-1,3 links.

They are synthesized naturally by a strain of the bacterium Leuconostoc mesenteroides.

Dextrans have been used extensively in the past for membrane characterization (Mehta

and Zydney, 2006; Mochizuki and Zydney, 1992) since they do not have any ionizable

side groups, thus providing a purely size-based measure of the membrane sieving

characteristics. The dextran diffusion coefficient is a function of the molecular weight as

evaluated by Granath (Granath, 1958):

[3.3]

where is the diffusion coefficient (in m2/s) and is the molecular weight in Da. The

Stokes radii can be evaluated from the diffusivity using the Stokes-Einstein equation to

give (Granath and Kvist, 1967):

35

[3.4]

with given in .

3.4 Ultrafiltration

3.4.1 Apparatus

Ultrafiltration experiments were performed in an Amicon 8010 stirred cell

(Millipore Corp., Bedford, MA). A membrane disc with effective area of 4.1 cm2 was

placed in the bottom of the stirred cell directly on top of a porous Tyvek® support that

provides membrane a mechanical support and minimizes deformation of the membrane at

high pressure. The stirred cell was placed on a magnetic stir plate with the stirring speed

set to 600 rpm using a Strobotac Type 1531-AB strobe light (General Radio Co.,

Concord, MA). An air-pressurized feed reservoir was connected to the stirred cell, with

the filtrate flux controlled by adjusting the pressure. A schematic of the apparatus is

shown in Figure 3.7.

36

Figure 3.7 Schematic of experimental set-up for constant pressure ultrafiltration

experiments.

3.4.2 Membrane Hydraulic permeability

The membrane hydraulic permeability ( ) was evaluated by measuring the

filtrate flux as a function of transmembrane pressure using a 10 mM buffered KCl

solution at pH 7. The permeability was evaluated from the slope of the data as:

[3.5]

where is the solution viscosity, is the filtrate flux, and is the transmembrane

pressure. Data were obtained at transmembrane pressures between 14 and 34 kPa (2 and

5 psi), with the filtrate flow rate evaluated by timed collection using a digital balance

(Model AG104, Mettler Toledo, Columbus, OH) with an accuracy of 100 μg.

37

3.4.3 Protein Sieving

Membrane sieving characteristics were evaluated using protein charge ladders to

directly study the effect of protein charge on the sieving coefficient. Each membrane was

first soaked in the charge ladder solution overnight at 4 °C to minimize initial transients

associated with protein adsorption on and within the membrane pores. The membrane

was then placed in the base of the stirred cell and the system was flushed with at least 25

L/m2 of buffer solution. The cell and feed reservoir were then filled with the charge

ladder solution, and the system was air-pressurized to approximately 10 kPa (1.5 psi).

Protein transmission was evaluated by collecting small samples of the filtrate and

bulk solutions after filtration of at least 500 μL to ensure equilibrium operation and to

clear the dead volume downstream of the membrane. Small samples of the bulk solution

were taken directly from the stirred cell (after clamping the filtrate port). All experiments

were performed at room temperature (22 ± 3 °C). The observed sieving coefficient was

calculated as:

[3.6]

where and are the protein concentrations in the filtrate and bulk solutions,

respectively. The stirred cell was carefully emptied, rinsed with DI water, and flushed

with at least 25 L/m2

of appropriate buffer between experiments. Dextran sieving

coefficients were evaluated using the same basic procedures, with the molecular weight

distribution of the dextrans in the feed and filtrate solutions analyzed using size exclusion

chromatography as discussed in Section 3.5.

38

3.4.4 Diafiltration

Diafiltration was performed using an Amicon 8200 stirred cell (Millipore

Corporation, Bedford, MA) with a 10 kDa Ultracel™ membrane to remove the dioxane

and any residual reactants formed during synthesis of the protein charge ladders. The

stirred cell was filled with the protein mixture and the feed reservoir was filled with

chilled DI water. The diafiltration was performed at a trans-membrane pressure of

approximately 6.9 kPa (1 psi) at room temperature (22 ± 3 °C) for at least four

diavolumes (defined as the ratio of the cumulative filtrate volume to the constant

retentate volume in the stirred cell).

3.4.5 Protein Fouling

Protein fouling experiments were performed with 5 g/L solutions of IgG. Limited

fouling experiments were also performed using 5 g/L solutions of α-lactalbumin,

lysozyme, and BSA. The membrane permeability was initially evaluated for each clean

membrane. The membrane was then soaked overnight in the protein solution, returned to

the stirred cell, rinsed with DI water, and the permeability re-evaluated. The membrane

was then flushed with at least 25 L/m2 of buffer at a pressure of 69 kPa (10 psi). The

stirred cell and feed reservoir were filled with the protein solution, the system was re-

pressurized to 69 kPa, and the filtrate flux was measured as a function of time for 60 min.

Filtrate samples were collected periodically for subsequent analysis, with bulk samples

collected directly from the stirred cell immediately before and after the experiment. After

39

completion of the ultrafiltration experiment, the stirred cell was carefully rinsed with DI

water and the buffer flux re-evaluated at 69 kPa. All experiments were performed at room

temperature (22 ± 3 °C).

The amount of protein adsorbed by the membrane in a static fouling system was

evaluated using the solution depletion method. A single membrane was placed in 10 mL

of a 5 g/L protein solution and allowed to soak overnight. The change in the protein

concentration in the solution was used to calculate the amount of protein adsorbed on the

membrane surface by a simple mass balance.

3.5 Size Exclusion Chromatography (SEC)

Dextran solutions were analyzed by size exclusion chromatography, also known

as gel permeation chromatography (GPC), to determine the concentration and molecular

weight distribution. An Agilent 1100 Series high performance liquid chromatography

system (Agilent Technologies, Palo Alto, CA) was used with a Superdex 200, 10/300

analytical column (13 μm particle size, 1 x 105 MW exclusion limit, from GE Healthcare,

Uppsala, Sweden). The mobile phase was a 10 mM Bis-Tris buffer at pH 7 containing

0.25 M KCl. The buffer was degassed before entering the system to avoid bubbles. The

column was initially equilibrated with a minimum of 2 column volumes of the mobile

phase at a flow rate of 0.3 mL/min. 25 μL samples of the dextran solution were then

injected by an autosampler immediately upstream of the guard column. The dextran

concentration in the exit stream was evaluated using a refractive index detector (Agilent

40

1100). Data collection was performed using ChemStation software version A.04.08

(Agilent Technologies, CA). Actual values were determined by comparison of the data

with that of known dextran standards.

3.6 Capillary electrophoresis

The concentration and net charge of each element of the charge ladders were

determined using a G1600A High-Performance Capillary Electrophoresis instrument

(Agilent Technologies, Palo Alto, CA) equipped with a dual polarity variable high

voltage DC power supply and variable wavelength UV-vis diode array detector.

Negatively charged fused silica capillaries (Agilent Technologies, G 1600-61211, Palo

Alto, CA) were used for the negatively charged α-lactalbumin, and positively charge

eCAP™ Amine capillaries (Beckman Coulter, Inc., 477431, Fullerton, CA) were used for

the positively charged lysozyme to minimize protein adsorption to the capillary wall.

Both capillaries had 50 μm inner diameters and were 65 cm in length (effective length of

approximately 55 cm). Protein detection was by UV absorbance at 214 nm. Mesityl oxide

(Fluka, 63940) was used as a neutral marker to evaluate the contribution of the electro-

osmotic flow. The capillaries were initially washed with 0.1 M NaOH for 10 min

followed by the running buffer (10 mM KCl in tris/glycine solution at pH 8.3) for an

additional 10 min. The eCAP™ Amine capillary was regenerated between runs using

amine regenerator solution (Beckman Coulter, Inc., 477433). 15-30 nL samples were

injected by application of a 3.5 kPa pressure for 25 s. Electrophoresis was performed at

an applied voltage around 25 kV. The electric field direction was chosen so that the

41

direction of the bulk flow was toward the outlet of the capillary. The current was kept

below 45 μA to minimize Joule heating. Electropherograms were recorded and analyzed

using 3D-CE ChemStation software (Version A.0903, Agilent Technologies, Palo Alto,

CA). Actual concentrations were evaluated by comparison with results for known protein

standards. Additional experimental details are available elsewhere (Ebersold and Zydney,

2004; Menon and Zydney, 1998).

3.7 X-ray Photoelectron Spectroscopy (XPS)

The extent of surface modification was estimated from the elemental composition

of the membrane as determined by X-ray photoelectron spectroscopy. The analysis was

performed using a Kratos Analytical Axis Ultra instrument (Kratos Analytical Inc.,

Chestnut Ridge, NY) available in the Materials Research Institute at The Pennsylvania

State University. The membrane was first flushed with deionized water, dried gently

using a Kimwipe, cut into small (approximately 12 mm x 5 mm) pieces using a razor

blade, and mounted on a sample platen. Data were obtained using monochromatic Al Kα

as the X-Ray source (1486.6 eV photons). The pressure in the analysis chamber was

Torr. The XPS data were analyzed using CasaXPS software (version 2.3.12Dev9)

by integrating the peak areas and applying the appropriate relative sensitivity factors to

account for the x-ray cross section and the transmission function of the spectrometer. All

binding energies were referred to the C1s peak at 285 eV.

42

Chapter 4

Fouling characteristics of zwitterionic membranes

Note: Most of the material presented in this Chapter was previously published (Hadidi

and Zydney, 2014)

4.1 Introduction

As discussed in Chapter 1, zwitterionic surfaces tend to adsorb very low amounts

of protein due to their high hydration capacity (via a combination of electrostatic and

hydrogen bonding interactions). The objective of the studies presented in this Chapter was

to obtain a fundamental understanding of the fouling behavior of a series of zwitterionic,

charged, and neutral membranes with nearly identical pore size / permeability. The

membranes were prepared by covalent attachment of small ligands to a base cellulose

(neutral) membrane, with the ligands specifically chosen to have the same effective size but

with different end functionality to generate an appropriate set of controls for the fouling

experiments.

4.2 Materials and Methods

The general procedures for the protein fouling experiments were described in

Chapter 3. Cellulose membranes were modified by activation with epichlorohydrin

43

followed by reaction with the appropriate ligand (Figures 3.3 and 3.4). Membranes were

characterized by both streaming potential measurements and XPS. Ultrafiltration

experiments were performed at constant transmembrane pressure using human serum

gamma globulin (IgG), bovine serum albumin (BSA), lysozyme, and α-lactalbumin as

model proteins. In addition, the amount of protein adsorbed on the membrane surface was

evaluated using the solution depletion method.

4.3 Results and Discussions

4.3.1 Membrane Modification

The extent of membrane modification was examined by X-ray photoelectron

spectroscopy (XPS). Figures 4.1 and 4.2 show XPS spectra for the upper surface

(approximately the upper 10 nm) of a zwitterionic, positive, acidic, hydrophilic,

hydrophobic, and sulfonic acid membrane. The spectra in Figure 4.1 were obtained

around the binding energy of nitrogen (398 eV). The nitrogen peak is absent in the base

cellulose (not shown) and in the membrane having the sulfonic acid modification but is

clearly visible in the other membranes. The double peak seen with the zwitterionic

membrane is likely associated with the presence of two distinct nitrogens, a primary

amine in the zwitterion and a secondary amine associated with the covalent linkage to the

membrane. Figure 4.2 shows the spectra around the sulfur peak (binding energy of 168

eV); in this case only the negatively charged membrane produced with the sulfonic acid

ligand showed a measurable peak in this range of binding energy as expected.

44

Figure 4.1 XPS spectra showing the nitrogen peak for all 100 kDa modified

membranes.

Figure 4.2 XPS spectra showing the sulfur peak for all 100 kDa modified membranes.

1000

1200

1400

1600

1800

2000

2200

2400

2600

385 390 395 400 405 410 415 420

Co

un

ts P

er

Seco

nd

Binding Energy (eV)

Hydrophobic

Positive

Acidic

Zwitterionic

Hydrophilic

Negative

100

200

300

400

500

600

700

150 155 160 165 170 175 180 185

Co

un

ts P

er

Seco

nd

Binding Energy (eV)

Negative

Acidic

Zwitterionic

Hydrophobic

hydrophilic

Positive

45

The atomic composition of the different membranes was determined from

quantitative analysis of the full XPS spectra based on a direct analysis of the peak area

associated with the C, O, N, and S atoms. The results are summarized in Table 4.1. The

nitrogen content was greatest for the zwitterionic and positively-charged membranes

since both of these membranes had two N in each ligand. The nitrogen content for the

acidic, hydrophobic, and hydrophilic membranes were similar, consistent with the

structure of these ligands (all with one N) and the use of the same epichlorohydrin

activation.

The degree of modification for each membrane was estimated from the atomic

fraction of nitrogen ( ) or sulfur as:

[4.1]

where f is the fraction of glucose rings in the base cellulose membrane that were

modified. The value of f is calculated assuming that all of the epichlorohydrin groups are

reacted with the ligand in the subsequent modification step. The parameters a, b, and d

are related to the atomic structure of the modified membrane: a is related to the number

of added N per ligand and b + df is the total number of atoms. For example, the

zwitterionic membrane modified with lysine has a=2 (accounting for two nitrogen atoms

in the two amine groups in lysine) and b=11 based on the number of carbon and oxygen

atoms in the base glucose ring (6 C and 5 O). The parameter d=14 based on the total

number of carbon, oxygen, and nitrogen atoms in the modified membrane (6 C, 2 O, and

2 N from the lysine along with 3 C and 1 O from the epichlorohydrin). This gives f =

0.025 for the zwitterionic membrane based on a nitrogen content of 0.43%.

46

The fraction of modified glucose rings for all membranes are summarized in

Table 4.1. All of the membranes prepared using the epichlorohydrin activation chemistry

had essentially the same degree of modification (f = 0.030 ± 0.004). The membrane

modified using the sulfonic acid ligand had a slightly lower degree of modification (f =

0.02) reflecting the difficulty in obtaining the same ligand density when using different

activation steps / reagents.

Table 4.1 Atomic composition (percent) and calculated degree of modification for

different 100 kDa modified membranes determined from XPS data.

Membrane f N Content C Content O Content S Content

Zwitterionic 0.025 ± 0.004 0.43 58.8 40.7 --

Positive 0.030 ± 0.004 0.52 60.0 39.4 --

Acidic 0.028 ± 0.004 0.25 60.5 39.2 --

Hydrophilic 0.026 ± 0.004 0.22 58.8 41.0 --

Hydrophobic 0.026 ± 0.004 0.23 60.6 39.1 --

Sulfonic acid 0.020 ± 0.004 -- 58.9 40.9 0.18

4.3.2 Membrane Surface Charge Characteristics

The effective surface charge of the different surface-modified ultrafiltration

membranes was evaluated from streaming potential measurements obtained with the fluid

flow directed through the membrane pores as described in Chapter 3. Figure 4.3 shows

47

typical experimental data for the measured streaming potential as a function of

transmembrane pressure ( ) for the zwitterionic version of the 100 kDa Ultracel™

membrane at different pH using 10 mM buffered KCl solution. The data were highly

linear at each pH with r2 values greater than 0.99. Repeat measurements for a given

membrane were highly reproducible with slopes within ±15%. The non-zero intercepts

are due to asymmetries in the Ag/AgCl electrodes and have no effect on the streaming

potential analysis (Burns and Zydney, 2000).

The data at pH 6, 7, and 8 have positive slopes, consistent with a net positive

charge on the membrane, while the slopes (and surface charge) are negative at pH 9 and

10. This behavior is due to deprotonation of the free amine functionality of the

zwitterionic membrane at high pH. The slope of the streaming potential versus pressure

data was used to calculate the apparent zeta potential of the membrane using the

Helmholtz-Smoluchowski equation (Equation 3.1):

[3.1]

with varying from +6.2 mV at pH 6 to -4.2 mV at pH 10.

48

Figure 4.3 Streaming potential data for 100 kDa Ultracel™ zwitterionic membrane in

10 mM buffered KCl solutions at several pH.

The charge characteristics of the zwitterionic membranes are examined in more

detail in Figure 4.4 which shows the measured values of the apparent zeta potential as a

function of the partial charge predicted using the pKa values for lysine: 9.06 for the

primary amine and 2.16 for the carboxylic acid (Brown, 1998). The pKa for the secondary

amine formed by the coupling of the lysine to the activated membrane was estimated as

8.8 based on the reported value for the primary amine in lysine (10.54) and the measured

difference in the pKa values of taurine (9.08) and TES (7.34) as reported by Rohani and

Zydney (Rohani and Zydney, 2012); the taurine and TES have analogous structures

except for the conversion of the primary amine in taurine into a secondary amine in TES.

The partial charge of the ligand was calculated using the Henderson-Hasselbach equation

as discussed in Chapter 2. The results are highly linear when plotted in this fashion with

r2 = 0.98, suggesting that the charge characteristics of the surface-grafted zwitterionic

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

5000 10000 15000 20000 25000 30000 35000

Str

eam

ing

po

ten

tial,

Ez (

mV

)

Transmembrane pressure, ΔP (Pa)

pH 6

pH 7

pH 8

pH 9

pH 10

49

membrane can be effectively described in terms of the protonation / deprotonation of the

acidic and basic components of the ligand.

Figure 4.4 Correlation between the apparent zeta potential and the calculated charge

based on the pKa values of the lysine ligand.

The results for the zwitterionic, positive, negative (sulfonic acid), hydrophilic, and

hydrophobic membranes at pH 7 are summarized in Table 4.2. The greatest positive

apparent zeta potential was obtained for the membrane modified with the diamine ligand

while the largest negative value was obtained with the sulfonic acid ligand as expected.

The zwitterionic (lysine-modified) membrane had a small positive charge due to the

secondary amine group formed by the covalent linkage to the cellulose membrane. This

secondary amine also provided the small positive charge on the hydrophilic and

hydrophobic membranes.

-5

-2.5

0

2.5

5

-1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1 1.25Ap

pare

nt

Zeta

Po

ten

tial,

𝜁𝑎𝑝𝑝 (m

V)

Calculated Fractional charge, Z

50

Table 4.2 The experimental values for the apparent zeta potential of surface-modified 100 kDa Ultracel™ membranes at pH 7 in 10 mM buffered KCl.

Membrane Zeta potential, (mV)

Zwitterionic 3.0 ± 0.1

Positively-charged 5.0 ± 0.1

Sulfonic Acid -9.4 ± 0.3

Hydrophilic 3.0 ± 0.1

Hydrophobic 4.3 ± 0.1

Acidic 0.5 ± 0.1

4.3.3 Dextran Ultrafiltration

To obtain additional insights into the sieving characteristics of the surface-

modified membranes, ultrafiltration experiments were performed using neutral dextrans.

The ultrafiltration was performed at relatively high ionic strength (150 mM) to minimize

electrostatic interactions. The dextran samples from the filtrate and feed were analyzed

by size exclusion chromatography, with the sieving coefficients calculated as the ratio of

the peak areas for narrow slices of the chromatogram (Mochizuki and Zydney, 1992).

Figure 5.5 shows results for the zwitterionic, hydrophilic, positive, and acidic membranes

over a wide range of dextran molecular weight. The sieving coefficients are plotted as a

function of the dextran radius, which was determined based on the retention time of

51

narrow molecular weight dextran standards using the correlation given by Equation [3.4].

The data for all four membranes are in fairly good agreement, indicating that these

membranes have very similar pore size characteristics. The dextran sieving coefficients

for the zwitterionic and hydrophilic membranes appear to be somewhat larger than those

for the acidic and positive membranes, which is consistent with the small differences in

hydraulic permeability for these membranes: for the zwitterionic and

hydrophilic compared to and for the acidic and positive

membranes, respectively.

Figure 4.5 Dextran sieving coefficients in 150 mM ionic strength at pH 7 using

different surface-modified membranes.

4.3.4 Static adsorption

The extent of static protein adsorption was evaluated using the solution depletion

method by measuring the change in protein concentration after incubation of the

membrane in a 5 g/L solution of serum IgG. Figure 5.1 shows results for the different

0

0.2

0.4

0.6

0.8

1

10 30 50 70 90

Sie

vin

g C

oeff

icie

nt,

So

Dextran Radius (Å)

Zwitterionic

Positive

Acidic

Hydrophilic

52

surface-modified membranes at pH 9 (i.e., where the IgG has a net negative charge),

plotted as the mass of protein adsorbed per unit cross-sectional area of the membrane.

The positively-charged membrane showed the greatest amount of protein adsorption,

consistent with the presence of an attractive electrostatic interaction. The zwitterionic and

negatively charged (sulfonic acid) membranes showed no measurable protein adsorption,

with the hydrophobic membrane showing approximately twice the amount of adsorption

as the hydrophilic surface.

Figure 4.6 Amount of protein adsorption from 5 g/L IgG at pH 9 on different modified

membranes.

Table 4.3 summarizes results for the membrane permeability before and after

overnight exposure to a 5 g/L serum IgG at different pH (in the absence of filtration). The

membranes used for these experiments were cut from adjacent areas of a single large

membrane sheet to ensure that the initial permeability of the unmodified membranes were

similar. The membrane permeability of the surface-modified membranes were also very

similar (variations of less than 10%), which is consistent with the use of the identical

0

200

400

600

800

1000

1200

Pro

tein

Ad

so

rpti

on

g/C

m2)

Zwitterionic

Hydrophilic

Hydrophobic

Positive

Sulfonic acid

53

activation chemistry coupled with the very similar size of the ligands. The permeability

of the negatively-charged membrane is slightly larger than that of the other membranes

due the smaller size of the sulfonic acid ligand (Figure 3.4). The permeability of the

zwitterionic membrane was largely unaffected by IgG adsorption, with less than a 12%

reduction in permeability under all conditions. The greatest fouling at pH 9 was seen with

the positively-charged membrane, consistent with the high amount of protein adsorption

due to the attractive electrostatic interactions under these conditions. The behavior at pH

5 was somewhat different since the IgG is positively-charged at this pH. Thus, the

negatively-charged membrane had the greatest reduction in permeability at pH 5.

Interestingly, the positively-charged membrane still showed more than a 30% reduction

in permeability at pH 5 even though the protein and membrane were like charged.

54

Table 4.3 Effect of protein adsorption on the permeability of the surface-modified 100 kDa Ultracel™ membranes.

Membrane

Clean membrane

Permeability, (m)

Pre-adsorbed

Permeability, (m)

%

Reduction

pH 5

Zwitterionic 3.6 0.4 3.4 0.3 6.9

Hydrophilic 3.8 0.4 3.6 0.4 5.1

Hydrophobic 3.5 0.4 3.1 0.3 13

Positive 3.5 0.4 2.4 0.2 32

Sulfonic Acid 3.8 0.4 2.0 0.2 48

pH 7

Zwitterionic 3.6 0.4 3.2 0.3 11

Hydrophilic 3.8 0.4 3.2 0.3 16

Hydrophobic 3.5 0.4 2.4 0.2 33

Positive 3.5 0.4 0.7 0.1 80

Sulfonic Acid 3.7 0.4 1.4 0.1 63

pH 9

Zwitterionic 3.3 0.3 3.0 0.3 11

Hydrophilic 3.3 0.3 2.0 0.2 39

Hydrophobic 3.3 0.3 1.8 0.2 47

Positive 3.3 0.3 1.3 0.1 61

Sulfonic Acid 3.5 0.4 2.3 0.2 30

4.3.5 IgG ultrafiltration

Protein ultrafiltration experiments were performed at a constant transmembrane

pressure of 69 kPa (10 psi). Figure 4.7 shows the flux data versus time for a single

experiment with a zwitterionic membrane, beginning with the buffer solution, followed

by the 5 g/L IgG, and then the buffer flux after the IgG filtration. The initial buffer flux

55

was 200 μm/s (720 L/m2/h), but this dropped to 11.8 μm/s with the start of the IgG

filtration due to the effects of concentration polarization and the osmotic pressure of the

highly retained IgG. The filtrate flux is nearly constant during the IgG ultrafiltration,

ranging from 11.8 to 11.2 µm/s, which suggests that there was relatively little fouling

over the course of the ultrafiltration process (beyond any immediate fouling upon

introduction of the protein solution at t = 10 min). At the end of the 1-hr filtration, the

stirred cell was then carefully emptied, refilled with buffer, and the flux re-evaluated at

the same transmembrane pressure. The buffer flux after the ultrafiltration was just

slightly greater than 160 µm/s, consistent with a small degree of fouling. Note that there

was no attempt made to clean the membrane at the end of the IgG filtration; the device

was simply emptied, refilled with a buffer solution, and then the stirrer was restarted.

Figure 4.7 Filtrate flux as a function of time during a typical fouling experiment for

the zwitterionic membrane using a 5 g/L IgG solution in 10 mM buffered

KCl at pH 7 and a constant pressure of 69 kPa (10 psi).

0

50

100

150

200

250

0 10 20 30 40 50 60 70 80

Flu

x (

µm

/s)

Time (min)

Buffer solution

Buffer solution

5 g/L IgG solution

56

The buffer flux and the initial filtrate flux with the 5 g/L IgG solution for the

different membranes are summarized in Table 4.4. The very low value of the buffer flux

for the positive membrane at pH 7 and 9 is due to protein adsorption; all of the

membranes examine in Table 4.4 were pre-soaked overnight in a 5 g/L solution of IgG at

the pH shown. The order of magnitude difference between the buffer and protein flux is

due to the large effect of concentration polarization. The zwitterionic membrane had the

highest protein flux at pH 7 and 9. The minimum in the flux for the zwitterionic

membrane occurred at pH 7, i.e., near the protein isoelectric point (pI), consistent with

previous experimental results (Fane et al., 1983). This is likely due primarily to the

increase in concentration polarization associated with the reduction in the protein

diffusion coefficient near the pI. The lowest flux at each pH was found with the

membrane that had a charge opposite to that of the protein. For example, the negatively-

charged membrane had the lowest protein flux at pH 5 (where the IgG was positively-

charged) while the positively-charged membrane had the lowest flux at pH 9.

57

Table 4.4 Buffer flux (after protein adsorption) and initial filtrate flux with a 5 g/L IgG solution for the different membranes at pH 5, 7, and 9.

pH 5 pH 7 pH 9

Buffer

flux,

( m/s)

Initial

protein

flux,

( m/s)

Buffer

flux,

( m/s)

Initial

protein

flux,

( m/s)

Buffer

flux,

( m/s)

Initial

protein

flux,

( m/s)

Zwitterionic 205 18.8 197 11.8 192 12.8

Hydrophilic 220 19.7 195 11.1 123 11.2

Positive 189 19.8 48 10.5 83.3 10.5

Sulfonic Acid 124 15.5 100 10.5 183 12.5

Figure 4.8 compares the measured values of the filtrate flux (top panel) and

filtrate concentration (bottom panel) during ultrafiltration of a 5 g/L IgG solution through

the different surface-modified membranes at pH 5. The flux for the zwitterionic and

positively-charged membranes are very similar, decreasing by approximately 15 % over

the course of the ultrafiltration. The flux data for the negatively-charged membrane are

quite different. The initial flux for the negative membrane was only 15.5 µm/s (compared

to 18.8 µm/s for the zwitterionic membrane), but the flux then increased slightly before

gradually decaying to a value that was very similar to that of the zwitterionic and

positively-charged membranes. The very different behavior for the negative membrane

may be related to a change in the properties of the pre-adsorbed protein on the surface of

this membrane at the start of the protein filtration. As seen in Tables 5.2 and 5.3, the

58

negatively-charged membrane had by far the greatest amount of protein adsorption at pH

5 (initial buffer flux of only 124 µm/s compared to 205 and 189 µm/s for the zwitterionic

and positively-charged membranes, respectively). Since the buffer permeability was

measured at pH 7, the adsorbed IgG would have been essentially neutral at the start of the

IgG filtration. Exposure of the pre-adsorbed IgG to the pH 5 buffer would lead to the

protonation of many of the ionizable amino acids, creating a net positive charge on the

protein and a corresponding expansion of the protein deposit and a reduction in the

hydraulic resistance to flow. This type of initial increase in flux was also seen in a fouling

experiment performed with the positively-charged membrane at pH 9, i.e., under

conditions where that membrane had a significant amount of pre-adsorbed protein.

Corresponding data for the observed sieving coefficients for IgG, defined as the

ratio of the filtrate to initial feed concentration, are shown in the bottom panel of Figure

4.8. The zwitterionic membrane had the largest initial sieving coefficient, which is

probably due to the smallest degree of protein adsorption. The IgG sieving coefficient for

the zwitterionic membrane remained relatively constant over the course of the

ultrafiltration; the slight decline in So is likely due to a combination of membrane fouling

and the small reduction in concentration polarization associated with the decrease in flux

over the course of the ultrafiltration. The greatest amount of IgG retention (smallest

sieving coefficients) was seen with the negatively-charged membrane, consistent with the

high degree of fouling seen with the negative membrane at pH 5, i.e., under conditions

where the protein and membrane are oppositely charged.

59

Figure 4.8 Filtrate flux (top panel) and filtrate concentration (bottom panel) for

ultrafiltration of a 5 g/L IgG solution in 10 mM buffered KCl at pH 5 at

a constant pressure of 69 kPa through the zwitterionic, positive, and

negative (sulfonic acid) membranes.

0

5

10

15

20

25

Flu

x (μ

m/s

)

Positive

Zwitterionic

Negative

0

0.02

0.04

0.06

0.08

0 10 20 30 40 50 60

Fil

trate

Co

ncen

trati

on

, C

f

Time

60

The results from the fouling experiments are summarized in Table 4.5 in terms of

the flux recovery ratio (F):

[4.2]

where and are the buffer flux of the clean membrane and the membrane after IgG

ultrafiltration, respectively. The flux recovery ratio (F) accounts for membrane fouling

under both static and dynamic flow conditions, with values of F close to one

corresponding to the absence of any significant fouling. The zwitterionic membrane had a

flux recovery ratio greater than 80% at all three pH, which is the best performance of any

of the membranes. This behavior is most pronounced at pH 9 where the zwitterionic

membrane had F = 0.82 while the best performance of the other membranes was F = 0.63

for the negatively charged membrane and only F = 0.36 for the positively charged

membrane. The hydrophilic membrane had very good performance at pH 5 (F = 0.84),

but this dropped to F = 0.57 at pH 9. The behavior of the positively- and negatively-

charged membranes were consistent with the differences in electrostatic interactions: the

performance of the negatively-charged membrane was best at pH 9 (F = 0.63), while the

performance of the positively-charged membrane was best at pH 5 (F = 0.58), in both

cases corresponding to conditions where the protein and membrane have the same

polarity (both negative or both positive, respectively).

61

Table 4.5 Flux recovery of different modified 100 kDa membranes after 1 hr ultrafiltration of an IgG solution at pH 5, 7, and 9.

pH Zwitterionic

Membrane

Hydrophilic

Membrane

Hydrophobic

Membranes

Positively-

charged

Membrane

Negatively-

charged

Membrane

5 89 84 63 58 45

7 80 74 57 26 35

9 82 57 49 36 63

Table 4.6 summarized the results for ultrafiltration of a 5 g/L IgG solution at pH 5

through the positive and zwitterionic membranes with and without the pre-adsorption

step. In both cases, the overall flux recovery was better for the membranes that were used

directly for the protein ultrafiltration, i.e., without exposing the membrane to IgG in the

pre-adsorption step. This effect was most pronounced for the positive membrane where

the flux recovery decreased from 72% to 58% when the membrane was pre-adsorbed

with IgG. The flux recovery for the zwitterionic membrane was 92% for the membrane

used directly for the protein ultrafiltration (without any pre-adsorption of the IgG).

62

Table 4.6 Initial buffer flux and flux recovery ratio for the zwitterionic and positively-

charged membranes after ultrafiltration of a 5 g/L IgG solution at pH 5 with

and without a pre-adsorption step.

Without adsorption With adsorption

Zwitterionic Positive Zwitterionic Positive

Initial Buffer Flux, (μm/s) 239 220 205 189

F (in %) 92 72 89 58

4.3.6 Protein ultrafiltration

Limited fouling experiments were also performed with other model proteins

having different size and surface charge characteristics to obtain additional insights into

the fouling resistance of the zwitterionic membranes. Experiments were performed

following the same basic procedures as used for the IgG ultrafiltration. This included

evaluating the clean membrane permeability, soaking the membranes overnight in a 5 g/L

solution of the protein, reevaluating the permeability, and then performing a constant

pressure ultrafiltration at 69 kPa for 1 hr. Table 4.7 summarizes the results in terms of the

flux recovery ratio, i.e., the ratio of the buffer flux after the protein ultrafiltration to that

of the clean membrane.

The zwitterionic membranes showed 92% and 81% flux recovery ratios after

ultrafiltration of lysozyme and α-lactalbumin at pH 7, respectively. These two proteins

have similar radius (1.6 nm, which is considerably smaller than the membrane pore size)

63

but very different surface charge characteristics, with the lysozyme being strongly

positively-charged while the α-lactalbumin has a significant negative charge at neutral pH

(Table 3.1). The lower flux recovery ratio after ultrafiltration of α-lactalbumin is

consistent with the weak attractive electrostatic interactions between the oppositely

charged α-lactalbumin and membrane at pH 7. The positively-charged membrane had a

similar flux recovery ratio as the zwitterionic membrane when used with lysozyme, but

the extent of irreversible fouling was much greater after ultrafiltration of the oppositely-

charged α-lactalbumin.

The data for BSA ultrafiltration were obtained at the protein isoelectric point (pH

4.7) and at relatively high solution ionic strength (150 mM), conditions that have

previously been shown to give high degrees of fouling (Fane et al., 1983). The

zwitterionic membrane showed F = 0.88 under these conditions compared to F = 0.76 for

the positively-charged membrane.

Table 4.7 Flux recovery of zwitterionic and positively-charged 100 kDa membranes

after 1 hr ultrafiltration of lysozyme and α-lactalbumin at pH 7 and BSA at pH 4.7.

Lysozyme α-Lactalbumin BSA

Zwitterionic 92 81 88

Positive 93 77 76

64

4.4 Conclusions

The data presented in this Chapter for the fouling behavior of the different surface

modified ultrafiltration membranes provide one of the most quantitative studies of the

performance of membranes with different surface functionalities (zwitterionic, positive,

negative, hydrophilic, and hydrophobic). The membranes were all prepared by covalent

modification of a base cellulose membrane with the same activation chemistry using a

series of ligands having essentially the same size/length but with different end-group

functionality. The resulting membranes had very similar hydraulic permeability and pore

size (dextran retention) but very different surface properties. This approach thus provides

an appropriate set of controls for understanding the effects of membrane surface

chemistry on the fouling characteristics.

The zwitterionic membrane displayed the greatest resistance to membrane fouling

over the full range of experimental conditions. This membrane showed negligible protein

adsorption in a static binding experiment, with the permeability of the protein-adsorbed

membrane decreasing only slightly compared to that of the pristine membrane.

Ultrafiltration of IgG caused a small additional reduction in permeability, but the flux

recovery for the zwitterionic membrane was greater than 80% at pH 5, 7, and 9

(corresponding to conditions where the protein was positively-charged, approximately

neutral, and negatively-charged). The hydrophilic membrane had a very similar flux

recovery value at pH 5, but this dropped to F < 0.6 when the IgG filtration was performed

at pH 9 while the zwitterionic membrane had F = 0.82 at this pH. The zwitterionic

membrane was also highly resistant to fouling of α-lactalbumin, lysozyme, and BSA,

65

with flux recovery ratios greater than 80% for all 3 proteins. This was true even for BSA

fouling at the protein isoelectric point, conditions that usually lead to extensive fouling.

Experimental data obtained with the positively- and negatively-charged

membranes clearly demonstrate that electrostatic interactions play an important role in

membrane fouling. The negatively-charged membrane showed its best performance

(greatest value of the flux recovery) at pH 9, where the IgG was also negatively charged.

In contrast, the positively-charged membrane had its best performance at pH 5 due to the

electrostatic repulsion between the protein and membrane under these conditions.

However, the largest flux recovery seen with the charged membranes was only F = 0.63,

which is well below the values obtained with the zwitterionic membrane.

Similarly, the results with the hydrophilic and hydrophobic membranes (generated

with the same ligands but with one having a hydroxyl group instead of a methyl group at

the end) show that hydrophilicity also has an impact on fouling, with the hydrophilic

membrane having a somewhat larger value of F over the entire pH range. However, the

flux recovery of the hydrophilic membrane remained below that for the zwitterionic

membrane at all three pH values. These results clearly demonstrate the very low fouling

characteristics of the zwitterionic surface, which could be very attractive for use in

ultrafiltration applications with highly fouling feed streams.

66

Chapter 5

Sieving characteristics of zwitterionic membranes

5.1 Introduction

As discussed in Chapter 1, it is now well-established that protein retention in

ultrafiltration is due to a combination of both protein size and electrostatic interactions

between the charged membrane and the charged protein. Rohani and Zydney (Rohani and

Zydney, 2012) recently extended these studies to zwitterionic membranes, performing

ultrafiltration experiments with acidic, basic, and neutral proteins. The protein sieving

coefficients were a strong function of solution conditions with the data for the different

proteins well correlated with the product of the surface charge densities of the protein and

membrane over a relatively wide range of conditions. However, the use of several different

proteins with different surface charge, charge distribution, and shape could easily have

obscured some of the key phenomena governing the behavior of these zwitterionic

membranes.

The primary objective of the studies presented in this Chapter was to use protein

charge ladders to quantitatively evaluate the effects of electrostatic interactions on protein

transport through a series of zwitterionic, charged, and neutral membranes prepared by

covalent attachment of small ligands to the base membrane. Protein charge ladders are

chemical derivatives of a protein differing by single charge groups, allowing the behavior of

67

a range of species with different charge, but essentially identical size / structure, to be

studied simultaneously in a single experiment.

5.2 Materials and Methods

The general procedures for evaluating the protein sieving characteristics were

described in Chapter 3. Membranes were modified by activation with epichlorohydrin

followed by reaction with the appropriate ligand (Figures 3.3 and 3.4). Membranes were

characterized by streaming potential measurements. Ultrafiltration experiments were

performed at constant transmembrane pressure using lysozyme and α-lactalbumin charge

ladders constructed by modification of the base protein as described in Chapter 3.

5.3 Results and Discussions

5.3.1 Membrane Characterization

As discussed in Chapters 3 and 4, the surface charge characteristics of the

different surface-modified ultrafiltration membranes were evaluated from streaming

potential measurements. The results for the zwitterionic, positively-charged, hydrophilic

and acidic membranes at pH 7 are summarized in Table 5.1. The greatest positive

apparent zeta potential was obtained for the membrane modified with the diamine ligand

(the positively-charged membrane). The zwitterionic (lysine-modified) and hydrophilic

68

membranes both had a small positive charge due to the protonation of the secondary

amine group formed by the covalent linkage to the cellulose membrane. The acidic

membrane was nearly neutral due to the positively-charged secondary amine and the

negatively charged carboxylic acid.

Table 5.1 Apparent zeta potential of surface-modified 100 kDa Ultracel™ membranes

at pH 7 in 10 mM buffered KCl.

Membrane Zeta potential, (mV)

Zwitterionic 3.0 ± 0.1

Positively-charged 5.0 ± 0.1

Hydrophilic 3.0 ± 0.1

Acidic 0.5 ± 0.1

5.3.2 Charge ladder Characterization

Figure 5.1 shows typical capillary electropherograms for 5 g/L solutions of the α-

lactalbumin (top panel) and lysozyme (bottom panel) charge ladders in a 10 mM ionic

strength tris/glycine buffer at pH 8.3. Transport in the capillary is dominated by the bulk

electroosmotic flow; the charged species migrate back against the electroosmotic flow

due to electrophoresis so that they pass the detector after the neutral marker. Thus, the

first peak in the electropherogram for α-lactalbumin represents the neutral marker

69

followed by the unmodified lactalbumin (which has the smallest negative charge) and

then the other protein derivatives, each of which has one less positive amine group (i.e.,

one more negative charge). 13 peaks or “rungs” are seen in the electropherogram,

corresponding to 0 to 12 acylated lysine groups. The electropherogram for the lysozyme

charge ladder begins with the most highly modified lysozyme species (with lowest net

positive charge at this pH), with the unmodified lysozyme eluting as the last peak. It is

just possible to make out 7 rungs in the charge ladder, consistent with the 6 lysine

residues in lysozyme.

70

Figure 5.1 Capillary electropherograms for α-lactalbumin (top panel) and lysozyme

(bottom panel) charge ladders in 10 mM tris/glycine buffer at pH 8.3.

The net charge for each of the protein variants was evaluated from the

electrophoretic mobility , which was determined from the migration times of the

neutral marker and the specific protein variant based on the location of the peak

maximum in the electropherogram as:

0

30

60

90

120

150

180

3.5 4.5 5.5 6.5 7.5 8.5 9.5 10.5

Ab

so

rban

ce (

mA

U)

Migration Time (min)

Neutral Marker

0

70

140

210

280

350

2.5 3 3.5 4 4.5 5

Ab

so

rban

ce (

mA

U)

Migration Time (min)

71

(

) [5.1]

where is the applied voltage, is effective capillary length, and and are the

migration times for the protein peak and the neutral marker, respectively. The net effective

charge was calculated from Equation 2.33 assuming that the protein is a uniformly charged

hard sphere (discussed in more detail in Chapter 2):

[2.33]

where is the protein radius, is the inverse Debye length (3.04 nm for the 10 mM buffer

solution used in the capillary electrophoresis), is the electron charge ,

and is Henry’s function (Henry, 1931) which accounts for the finite double layer

thickness (given by Equation 2.32).

The open symbols in Figure 5.2 show the calculated charge of each rung in the

charge ladder as determined from the capillary electrophoresis using Equations 5.1 and

2.33. Data for lysozyme were obtained using the positively-charged eCAP™ Amine

capillaries while those for α-lactalbumin were obtained with the negatively-charged fused

silica capillaries to minimize protein adsorption. The filled symbols in Figure 5.2 show

the calculated values of the net protein charge for the corresponding rungs calculated

from the amino acid composition by eliminating one lysine group for each rung. The

native lysozyme appears as peak 7 while the native α-lactalbumin is peak 1. The

calculated values of the lysozyme charge are in good agreement with the values from the

capillary electrophoresis; the deviations are somewhat larger for α-lactalbumin. The small

difference between the model calculations and the CE results could simply be due to the

72

approximations in the analysis, e.g., the assumption of a uniformly charged sphere;

although this might also reflect a small amount of ion binding in the CE experiments as

well as the effects of charge regulation (discussed below).

The net protein charge varies almost linearly with peak number, although the

change in charge between peaks is smaller than the one charge unit that would be

expected for the removal of one amine. For example, the net charge on α-lactalbumin

from CE decreases by 4.4 electronic charges as one goes from the unmodified protein

(Z=-4.3) to the protein with 6 reacted amine groups (Z=-8.7), which is less than the

expected reduction of 6 electronic charges. This behavior is due to charge regulation

effects associated with the change in protonation of the other charged amino acid residues

associated with the alteration in the local H+ concentration at the surface of the protein

caused by the change in net protein charge. For example, the increase in net negative

charge caused by the removal of one amine causes an increase in the local H+

concentration and a corresponding shift in the acid-base equilibrium leading to a small

increase in the degree of protonation. This is discussed in more detail elsewhere (Menon

and Zydney, 2000).

73

Figure 5.2 Net charge for the first seven peaks in the α-lactalbumin and lysozyme

charge ladders evaluated both from the electrophoretic mobility data

using 10 mM tris/glycine buffer at pH 8.3 (open symbols) and from the

amino acid composition (filled symbols).

5.3.3 Ultrafiltration Experiments

5.3.3.1 Protein Ultrafiltration

Figure 5.3 shows typical results for the observed sieving coefficients of several

“rungs” of the lysozyme charge ladder (i.e., individual peaks in the capillary

electropherogram) for the zwitterionic, hydrophilic, and positively-charged membranes.

The data are plotted as a function of the net protein charge determined from the amino

acid sequence; it was not possible to evaluate the charge at pH 7 from the electrophoretic

mobility due to the poor resolution of the capillary electrophoresis at this pH. The

-12

-10

-8

-6

-4

-2

0

2

4

6

8

0 1 2 3 4 5 6 7 8

Net

pro

tein

ch

arg

e, Z

Peak Number

Lysozyme-CE

α-lactalbumin-CE

α-lactalbumin-Amino acid

Lysozyem-Amino acid

74

observed sieving coefficient is defined as the ratio of the protein concentration in the

filtrate solution to that in the feed, with the concentration of each rung of the charge

ladder evaluated directly from the capillary electropherogram by numerical integration

under the peak (defined between adjacent local minima in the electropherogram). The

results for each membrane thus represent data obtained in a single ultrafiltration

experiment. The sieving coefficients for the zwitterionic, positively-charged, and

hydrophilic membranes decrease with increasing protein charge, which is a direct result

of the electrostatic repulsion between the positively-charged membranes and the

positively-charged lysozyme. The sieving coefficients for the zwitterionic and

hydrophilic membranes were very similar, consistent with the similar zeta potential for

these membranes, suggesting that the zwitterionic functionality at the end of the ligand

acts like a hydrophilic group. The sieving coefficients for the positively-charged

membrane display a greater slope, i.e., a greater reduction in sieving coefficient with

increasing protein charge, along with the a lower apparent intercept at zero protein

charge. This behavior is likely due to the larger positive charge (+5 mV compared to +3

mV for the zwitterionic and hydrophilic membranes) as well as the somewhat smaller

initial permeability (3.3 x 10-12

m compared to 3.6 x 10-12

m for the zwitterionic and

hydrophilic membranes), with the latter corresponding to a smaller effective pore size.

This smaller pore size is also seen in the slightly smaller values of the dextran sieving

coefficients for the positively-charged membrane (Figure 4.5).

75

Figure 5.3 Observed sieving coefficients for ultrafiltration of lysozyme charge ladder

at pH 7 through 100 kDa modified Ultracel™ membrane as a function of

net protein charge.

In order to obtain additional insights into the electrostatic interactions, the

observed sieving coefficient data for the different membranes are plotted in Figure 5.4 as

a function of the product of the dimensionless surface charge densities of the protein and

the membrane. This form is consistent with the theoretical analysis of the partitioning of

a charged sphere in a charged cylindrical pore (Smith and Deen, 1980) assuming that the

electrostatic interactions are dominated by the term arising from the interaction between

the electrical double layers of the protein and pore. The surface charge density for each

variant was calculated from the amino acid composition while the dimensionless

membrane surface charge density was evaluated from the apparent zeta potential

measurement at pH 7. The solid curves represent the theoretical calculations for

membranes with effective pore radii of 7.2 nm (blue and orange curves) and 9.2 nm (red

0

0.2

0.4

0.6

0.8

1

0 1 2 3 4 5 6 7 8

Ob

serv

ed

Sie

vin

g C

oeff

icie

nt,

So

Net Protein Charge, Z

Zwitterionic

Hydrophilic

Positive

76

curve), determined by minimizing the sum of the squared residuals between the model

and data for the positively-charged / acidic membranes and for the zwitterionic /

hydrophilic membranes, respectively. The sieving coefficients decrease monotonically

with increasing values of the charge interaction parameter as expected. The small

difference between the blue and orange curves is due to the terms involving the square of

the protein charge and membrane charge in Equation 2.12. These terms are small relative

to the term involving the product of the protein and membrane charge, although they do

have a measurable effect on the sieving coefficient at large values of the protein charge.

The significant difference in the best fit values of the pore radius for the

positively-charged / acidic membranes and the zwitterionic / hydrophilic membranes is

surprising. The positively-charged / acidic membranes did have a slightly lower hydraulic

permeability, although the approximately 10% difference in permeabilities would

correspond to only a 5% difference in pore radius (assuming Poiseuille flow). One

possible explanation for this effect is the different structure of the charged ligands used to

construct the surface-modified membranes. The positively-charged and acidic

membranes both have charged groups at the far end of the ligand, towards the center of

the pore, and thus in close proximity to the protein. In contrast, the zwitterionic and

hydrophilic membranes have end groups that are electrically neutral, with the small

positive charge on these membranes arising from the secondary amine that is located

closer to the pore wall (and thus further away from the protein). Rohani et al. (Rohani et

al., 2010) reported the size of a diaminodecane ligand as 1.4 nm, which has the same

physical structure (but slightly larger size) as the diaminohexane ligand used to generate

the positively-charged membrane in this thesis. This is still somewhat smaller than the 2

77

nm difference in the best fit values of the pore size, but it is at least qualitatively

consistent with the experimental observations.

Figure 5.4 Observed sieving coefficients as a function of charge interaction parameter

(product of the dimensionless surface charge densities for the protein and

the membrane) for different membranes. Filled symbols represent the

experimental data with the solid curves representing the theoretical model.

5.4 Conclusions

Although previous studies of protein transport through semipermeable

ultrafiltration membranes clearly demonstrated the importance of electrostatic

interactions, there have been no quantitative studies comparing the sieving characteristics

of zwitterionic membranes with those of charged and neutral membranes with otherwise

78

similar properties. The experimental studies performed in this Chapter used protein

charge ladders consisting of a series of protein variants with the same 3-dimensional

structure and size but with different surface charge to probe the sieving characteristics of

the different membranes, allowing data to be taken over a range of protein charge in a

single experiment. This eliminates artifacts associated with differences in protein fouling

or variability in membrane properties between experiments.

Protein transmission through the zwitterionic, positively-charged, acidic, and

hydrophilic membranes was highly correlated with the product of the protein and

membrane charge, consistent with available theoretical models based on the partitioning

of a charged sphere in a charged cylindrical pore. The sieving coefficients for the

zwitterionic membrane were nearly identical to those for the hydrophilic membrane,

indicating that the zwitterionic group behaves like an uncharged hydrophilic (hydroxyl)

group in the context of protein transport. Interestingly, the model calculations suggest

that the zwitterionic and hydrophilic membranes have an effective pore size that is 2 nm

larger than that of the positively-charged and acidic membranes. This discrepancy

appears to be due to the different charge structure of the ligands, with the charge groups

in the positive and acidic membranes located at the end of the ligand (facing out into the

center of the pore) while the small positive charge on the zwitterionic and hydrophilic

membranes is associated with the secondary amine located much closer to the cellulose

surface. This suggests that these zwitterionic membranes might be able to provide the

enhanced permeability – selectivity characteristic of electrically charged membranes

while maintaining very low fouling characteristics due to the zwitterion at the outer

79

portion of the ligand. Additional experimental studies will be needed to confirm these

experimental results and provide further support for this physical picture.

80

Conclusion and future work

6.1 Introduction

The production of high value recombinant proteins requires robust, cost-effective,

and high-resolution separation methods that can provide high yield and purification of the

desired product. Ultrafiltration processes have remarkable potential to meet these needs

since they provide high throughput protein purification under mild conditions that will

not degrade or damage the biological product. One of the major challenges in many UF

processes is membrane fouling due to interactions between the proteins and membrane

surface. Several recent studies have shown that zwitterionic surfaces have low protein

adsorption characteristics due to their high degree of hydration associated with electrostatic

and hydrogen bonding interactions.

This thesis provides a quantitative study of the transport and fouling characteristics

of zwitterionic membranes in comparison with electrically charged and neutral membranes.

Data were obtained for a series of membranes prepared by covalent modification of a base

cellulose membrane using the same activation chemistry with a series of ligands having

essentially the same size/length but different end-group functionality. This approach thus

provided an appropriate set of controls for understanding the effect of membrane surface

chemistry on both fouling and protein transport. The following subsections summarize the

key experimental and theoretical results from the different parts of this thesis.

Recommendations for future work are also discussed.

81

6.2 Protein Transport through Surface Modified Membranes

Although previous studies of protein transport through semipermeable

zwitterionic membranes clearly demonstrated the importance of electrostatic interactions,

those data were obtained at different pH and with proteins having different surface charge

characteristics. The experimental studies performed in this thesis used protein charge

ladders consisting of a series of protein variants with the same 3-dimensional structure

and size but with different surface charge. Protein transmission though the zwitterionic,

positively-charged, acidic, and hydrophilic membranes was highly correlated with the

product of the protein and membrane charge, consistent with predictions of available

theoretical models based on the partitioning of a charged sphere in a charged cylindrical

pore. The very similar behavior of the zwitterionic and hydrophilic membranes suggests

that the zwitterionic ligand behaves as a hydrophilic (uncharged) functionality.

Interestingly, the model calculations indicate that the zwitterionic and hydrophilic

membranes have a much smaller effective pore size than the positive and acidic

membranes, even though all four membranes have fairly similar pore size based on the

measured values of the hydraulic permeability and dextran sieving coefficients. This is

likely due to the different structure of the ligands, with the positive charge on the

zwitterionic and hydrophilic membranes arising from the secondary amine that is located

near the polymer surface and thus a significant distance away from the protein within the

membrane pore.

82

6.3 Protein Fouling of Surface Modified Membranes

The data presented in Chapter 4 provided one of the most quantitative studies of

the fouling behavior of membranes with different surface functionalities (zwitterionic,

positive, negative, hydrophilic, and hydrophobic). The zwitterionic membrane displayed

the greatest resistance to membrane fouling over the full range of experimental

conditions. This membrane showed negligible protein adsorption in a static binding

experiment, with the permeability of the protein-adsorbed membrane decreasing only

slightly compared to that of the pristine membrane. Ultrafiltration of IgG caused a small

additional reduction in permeability, but the flux recovery for the zwitterionic membrane

was greater than 80% at pH 5, 7, and 9 (corresponding to conditions where the protein

was positively-charged, approximately neutral, and negatively-charged).

The data with the positively- and negatively-charged membranes clearly

demonstrated that electrostatic interactions play an important role in membrane fouling,

with the greatest fouling seen when the membrane and protein are oppositely charged.

Surface hydrophilicity also plays a role, with the hydrophilic membrane having a better

flux recovery than the corresponding hydrophobic membrane. However, the largest flux

recovery seen with the charged membranes was only F = 0.63, which is well below the

values obtained with the zwitterionic membrane. The very low fouling characteristics of

the zwitterionic surface make this membrane very attractive for use in protein

ultrafiltration.

83

6.4 Recommendations

The results presented in this thesis provide important insights into the protein

transport and fouling behavior of the zwitterionic membranes. However, there are a

number of important areas that would benefit from additional experimental and theoretical

investigations.

The experimental studies presented in this thesis were performed using a small

stirred cell. However, industrial ultrafiltration systems use tangential flow filtration (TFF)

modules that have much better mass transfer characteristics. The filtrate flux in these TFF

modules is likely to be more strongly influenced by membrane fouling due to the reduction

in concentration polarization effects; future experimental studies should be performed to

directly evaluate the flux and flux recovery for protein ultrafiltration in these tangential

flow modules. These studies should also examine the stability of the zwitterionic surfaces

upon exposure to strong cleaning solutions (like NaOH and NaOCl) and after multiple

ultrafiltration cycles.

It would also be very desirable to extend this work to other ultrafiltration

applications that require low fouling membranes. For example, ultrafiltration is used

extensively in water treatment applications, providing significant removal of natural

organic matter and viruses. The zwitterionic membranes produced in this thesis could be

very attractive in these water treatment applications if they are able to retain their very low

fouling characteristics. Other systems of interest would include DNA, surfactants, and

harvested cell culture fluid, all of which tend to be difficult to process using ultrafiltration

due to the high degree of fouling.

84

It would also be interesting to explore the further optimization of the ligand

structure for these zwitterionic UF membranes. For example, the zwitterionic ligand

could be attached to the membrane using a different linker, e.g., using a chemistry that

generates a fully neutral surface instead of the positively-charged surface produced with

the amine linkage examined in this thesis. It would also be possible to tune the properties

of the linker or the zwitterionic ligand, e.g., by optimizing the distance between the two

charged groups in the zwitterion or the length of the connector between the zwitterion and

the base membrane. It would also be interesting to perform experiments with “mixed-

charge” membranes generated by covalent attachment of a combination of separate

negatively-charged and positively-charged ligands to the base membrane. This would make

it possible to explore the effect of the charge distribution over the membrane surface on the

performance of the zwitterionic membrane.

The low fouling characteristics of zwitterionic surfaces are typically attributed to

the high degree of hydration of the zwitterion. More fundamental insights into the effects of

water dynamics, ligand chemistry/structure, and ligand distribution could potentially be

obtained using appropriate molecular dynamics simulations to evaluate the energy of

interaction between the protein and membrane surface. These simulations could then be

used to aid the design and selection of zwitterionic ligands for very low fouling

membranes.

85

References

An, Q.-F., Sun, W.-D., Zhao, Q., Ji, Y.-L., and Gao, C.-J. (2013). Study on a novel

nanofiltration membrane prepared by interfacial polymerization with zwitterionic

amine monomers. Journal of Membrane Science 431, 171-179.

Andersen, S., Blouch, K., Bialek, J., Deckert, M., Parving, H.-H., and Myers, B. D.

(2000). Glomerular permselectivity in early stages of overt diabetic nephropathy.

Kidney international 58, 2129-2137.

Anderson, J. L., and Quinn, J. A. (1974). Restricted transport in small pores: a model for

steric exclusion and hindered particle motion. Biophysical Journal 14, 130-150.

Brown, T. A. (1998). "Molecular Biology Labfax: Recombinant DNA," Academic Pr.

Bungay, P. M., and Brenner, H. (1973). The motion of a closely-fitting sphere in a fluid-

filled tube. International Journal of Multiphase Flow 1, 25-56.

Burns, D. B. (2000). Electrostatic interactions in protein trasnport using ultrafitration

membranes: Effects of solution conditions and membrane charges, University of

Delaware, Newark, DE.

Burns, D. B., and Zydney, A. L. (2000). Buffer effects on the zeta potential of

ultrafiltration membranes. Journal of membrane science 172, 39-48.

Chang, R., Deen, W. M., Robertson, C. R., and Brenner, B. M. (1975). Permselectivity of

the glomerular capillary wall: III. Restricted transport of polyanions. Kidney int 8,

212-218.

Chelmowski, ., K ster, S. D., Kerstan, A., Prekelt, A., Grunwald, C., Winkler, T.,

Metzler-Nolte, N., Terfort, A., and W ll, C. (2008). Peptide-based SAMs that

resist the adsorption of proteins. Journal of the American Chemical Society 130,

14952-14953.

Chen, S., Zheng, J., Li, L., and Jiang, S. (2005). Strong resistance of phosphorylcholine

self-assembled monolayers to protein adsorption: insights into nonfouling

properties of zwitterionic materials. Journal of the American Chemical Society

127, 14473-14478.

86

Chung, W. K., Evans, S. T., Freed, A. S., Keba, J. J., Baer, Z. C., Rege, K., and Cramer,

S. M. (2009). Utilization of lysozyme charge ladders to examine the effects of

protein surface charge distribution on binding affinity in ion exchange systems.

Langmuir 26, 759-768.

Clark, D. S., and Blanch, H. W. (1997). "Biochemical Engineering, Second Edition,"

Taylor & Francis.

Deen, W. (1987). Hindered transport of large molecules in liquid‐filled pores. AIChE

Journal 33, 1409-1425.

Ebersold, M. F., and Zydney, A. L. (2004). Use of protein charge ladders to study

electrostatic interactions during protein ultrafiltration. Biotechnology and

bioengineering 85, 166-176.

Fane, A., Fell, C., and Suki, A. (1983). The effect of pH and ionic environment on the

ultrafiltration of protein solutions with retentive membranes. Journal of

Membrane Science 16, 195-210.

Granath, K. A. (1958). Solution properties of branched dextrans. Journal of Colloid

Science 13, 308-328.

Granath, K. A., and Kvist, B. E. (1967). Molecular weight distribution analysis by gel

chromatography on Sephadex. Journal of Chromatography A 28, 69-81.

Guo, W., Ngo, H.-H., and Li, J. (2012). A mini-review on membrane fouling.

Bioresource Technology.

Hadidi, M., and Zydney, A. L. (2014). Fouling behavior of zwitterionic membranes:

Impact of electrostatic and hydrophobic interactions. Journal of Membrane

Science 452, 97-103.

Harder, P., Grunze, M., Dahint, R., Whitesides, G., and Laibinis, P. (1998). Molecular

conformation in oligo (ethylene glycol)-terminated self-assembled monolayers on

gold and silver surfaces determines their ability to resist protein adsorption. The

Journal of Physical Chemistry B 102, 426-436.

87

He, Y., Hower, J., Chen, S., Bernards, M. T., Chang, Y., and Jiang, S. (2008). Molecular

simulation studies of protein interactions with zwitterionic phosphorylcholine

self-assembled monolayers in the presence of water. Langmuir 24, 10358-10364.

Henry, D. (1931). The cataphoresis of suspended particles. Part I. The equation of

cataphoresis. Proceedings of the Royal Society of London. Series A 133, 106-129.

Hunter, R. J. (1981). "Zeta potential in colloid science: principles and applications,"

Academic press London.

Ishihara, K., Nomura, H., Mihara, T., Kurita, K., Iwasaki, Y., and Nakabayashi, N.

(1998). Why do phospholipid polymers reduce protein adsorption? Journal of

biomedical materials research 39, 323-330.

Ishihara, K., and Takai, M. (2009). Bioinspired interface for nanobiodevices based on

phospholipid polymer chemistry. Journal of The Royal Society Interface 6, S279-

S291.

Ji, Y.-L., An, Q.-F., Zhao, Q., Sun, W.-D., Lee, K.-R., Chen, H.-L., and Gao, C.-J.

(2012). Novel composite nanofiltration membranes containing zwitterions with

high permeate flux and improved anti-fouling performance. Journal of Membrane

Science 390, 243-253.

Lebreton, B., Brown, A., and van Reis, R. (2008). Application of high‐performance

tangential flow filtration (HPTFF) to the purification of a human pharmaceutical

antibody fragment expressed in Escherichia coli. Biotechnology and

Bioengineering 100, 964-974.

Lewis, A. L. (2000). Phosphorylcholine-based polymers and their use in the prevention of

biofouling. Colloids and Surfaces B: Biointerfaces 18, 261-275.

Li, J.-H., Li, M.-Z., Miao, J., Wang, J.-B., Shao, X.-S., and Zhang, Q.-Q. (2012a).

Improved surface property of PVDF membrane with amphiphilic zwitterionic

copolymer as membrane additive. Applied Surface Science 258, 6398-6405.

Li, M.-Z., Li, J.-H., Shao, X.-S., Miao, J., Wang, J.-B., Zhang, Q.-Q., and Xu, X.-P.

(2012b). Grafting zwitterionic brush on the surface of PVDF membrane using

physisorbed free radical grafting technique. Journal of membrane science.

88

Li, Q., Bi, Q.-Y., Zhou, B., and Wang, X.-L. (2012c). Zwitterionic sulfobetaine-grafted

poly (vinylidene fluoride) membrane surface with stably anti-protein-fouling

performance via a two-step surface polymerization. Applied Surface Science 258,

4707-4717.

Liu, Y.-C., Suen, S.-Y., Huang, C.-W., and ChangChien, C.-C. (2005). Effects of spacer

arm on penicillin G acylase purification using immobilized metal affinity

membranes. Journal of membrane science 251, 201-207.

Malone, D. M., and Anderson, J. L. (1978). Hindered diffusion of particles through small

pores. Chemical Engineering Science 33, 1429-1440.

Mehta, A. (2006). Performance characteristics of charged ultrafiltration membranes:

fundamental studies and applications, The Pennsylvania State University,

University Park, PA.

Mehta, A., and Zydney, A. L. (2006). Effect of membrane charge on flow and protein

transport during ultrafiltration. Biotechnology progress 22, 484-492.

Mehta, A., and Zydney, A. L. (2008). Effect of spacer arm length on the performance of

charge-modified ultrafiltration membranes. Journal of Membrane Science 313,

304-314.

Menon, M. K. (1999). Effects of protein charge and electrostatic interactions in

ultrafiltration and size exclusion chromatography, University of Delaware,

Newark, DE.

Menon, M. K., and Zydney, A. L. (1998). Measurement of protein charge and ion binding

using capillary electrophoresis. Analytical chemistry 70, 1581-1584.

Menon, M. K., and Zydney, A. L. (2000). Determination of effective protein charge by

capillary electrophoresis: effects of charge regulation in the analysis of charge

ladders. Analytical chemistry 72, 5714-5717.

Mochizuki, S., and Zydney, A. L. (1992). Dextran transport through asymmetric

ultrafiltration membranes: comparison with hydrodynamic models. Journal of

membrane science 68, 21-41.

89

Mochizuki, S., and Zydney, A. L. (1993). Theoretical analysis of pore size distribution

effects on membrane transport. Journal of membrane science 82, 211-227.

Molek, J. R. (2008). Ultrafiltration of PEGylated Proteins, The Pennsylvania State

University, University Park, PA.

Molek, J. ., and Zydney, A. L. (2007). Separation of PEGylated α‐Lactalbumin from

Unreacted Precursors and Byproducts Using Ultrafiltration. Biotechnology

progress 23, 1417-1424.

Nagumo, R., Akamatsu, K., Miura, R., Suzuki, A., Tsuboi, H., Hatakeyama, N., Takaba,

H., and Miyamoto, A. (2012). Assessment of the Antifouling Properties of

Polyzwitterions from Free Energy Calculations by Molecular Dynamics

Simulations. Industrial & Engineering Chemistry Research 51, 4458-4462.

Opong, W. S., and Zydney, A. L. (1991). Diffusive and convective protein transport

through asymmetric membranes. AIChE journal 37, 1497-1510.

Overbeek, J. T. G., and Wiersema, P. H. (1967). The interpretation of electrophoretic

mobilities. Electrophoresis 2, 1-52.

Pasmore, M., Todd, P., Smith, S., Baker, D., Silverstein, J., Coons, D., and Bowman, C.

N. (2001). Effects of ultrafiltration membrane surface properties on< i>

Pseudomonas aeruginosa</i> biofilm initiation for the purpose of reducing

biofouling. Journal of Membrane Science 194, 15-32.

Pujar, N. S., and Zydney, A. L. (1994). Electrostatic and electrokinetic interactions

during protein transport through narrow pore membranes. Industrial &

engineering chemistry research 33, 2473-2482.

Pujar, N. S., and Zydney, A. L. (1997). Charge regulation and electrostatic interactions

for a spherical particle in a cylindrical pore. Journal of Colloid and Interface

Science 192, 338-349.

Razi, F., Sawada, I., Ohmukai, Y., Maruyama, T., and Matsuyama, H. (2012). The

improvement of antibiofouling efficiency of polyethersulfone membrane by

functionalization with zwitterionic monomers. Journal of Membrane Science 401,

292-299.

90

Rohani, M. M., Mehta, A., and Zydney, A. L. (2010). Development of high performance

charged ligands to control protein transport through charge-modified

ultrafiltration membranes. Journal of Membrane Science 362, 434-443.

Rohani, M. M., and Zydney, A. L. (2009). Effect of surface charge distribution on protein

transport through semipermeable ultrafiltration membranes. Journal of Membrane

Science 337, 324-331.

Rohani, M. M., and Zydney, A. L. (2012). Protein transport through zwitterionic

ultrafiltration membranes. Journal of Membrane Science 397, 1-8.

Saksena, S., and Zydney, A. L. (1994). Effect of solution pH and ionic strength on the

separation of albumin from immunoglobulins (IgG) by selective filtration.

Biotechnology and bioengineering 43, 960-968.

Saksena, S., and Zydney, A. L. (1995). Pore size distribution effects on electrokinetic

phenomena in semipermeable membranes. Journal of membrane science 105,

203-215.

Sharma, U., Negin, R. S., and Carbeck, J. D. (2003). Effects of cooperativity in proton

binding on the net charge of proteins in charge ladders. The Journal of Physical

Chemistry B 107, 4653-4666.

Shi, Q., Su, Y., Chen, W., Peng, J., Nie, L., Zhang, L., and Jiang, Z. (2011). Grafting

short-chain amino acids onto membrane surfaces to resist protein fouling. Journal

of Membrane Science 366, 398-404.

Smith, F. G., and Deen, W. M. (1980). Electrostatic double-layer interactions for

spherical colloids in cylindrical pores. Journal of Colloid and Interface Science

78, 444-465.

Smith, F. G., and Deen, W. M. (1983). Electrostatic effects on the partitioning of

spherical colloids between dilute bulk solution and cylindrical pores. Journal of

Colloid and Interface Science 91, 571-590.

Smith, K., Colton, C., Merrill, E., and Evans, L. (1968). Convective transport in a batch

dialyzer: determination of true membrane permeability from a single

measurement. In "Chem. Eng. Prog. Symp. Ser", Vol. 64, pp. 45.

91

Sun, Q., Su, Y., Ma, X., Wang, Y., and Jiang, Z. (2006). Improved antifouling property

of zwitterionic ultrafiltration membrane composed of acrylonitrile and

sulfobetaine copolymer. Journal of membrane science 285, 299-305.

Thömmes, J., and Etzel, M. (2007). Alternatives to chromatographic separations.

Biotechnology Progress 23, 42-45.

van Eijndhoven, R. H., Saksena, S., and Zydney, A. L. (1995). Protein fractionation using

electrostatic interactions in membranel filtration. Biotechnology and

Bioengineering 48, 406-414.

Van Reis, R., Brake, J., Charkoudian, J., Burns, D., and Zydney, A. (1999). High-

performance tangential flow filtration using charged membranes. Journal of

Membrane Science 159, 133-142.

van Reis, R., and Zydney, A. (2007). Bioprocess membrane technology. Journal of

Membrane Science 297, 16-50.

Van Reis, R. D. (2006). Charged filtration membranes and uses therefor. Google Patents.

Vermette, P., and Meagher, L. (2003). Interactions of phospholipid-and poly (ethylene

glycol)-modified surfaces with biological systems: relation to physico-chemical

properties and mechanisms. Colloids and Surfaces B: Biointerfaces 28, 153-198.

Vrijenhoek, E. M., Hong, S., and Elimelech, M. (2001). Influence of membrane surface

properties on initial rate of colloidal fouling of reverse osmosis and nanofiltration

membranes. Journal of Membrane Science 188, 115-128.

Wang, T., Wang, Y.-Q., Su, Y.-L., and Jiang, Z.-Y. (2006). Antifouling ultrafiltration

membrane composed of polyethersulfone and sulfobetaine copolymer. Journal of

membrane science 280, 343-350.

Young, M., Carroad, P., and Bell, R. (1980). Estimation of diffusion coefficients of

proteins. Biotechnology and Bioengineering 22, 947-955.

Zeman, L., and Zydney, A. (1996). Microfiltration and Ultrafiltration: Principles and

Applications. Marcel Dekker, New York.

92

Zydney, A. L. (1997). Stagnant film model for concentration polarization in membrane

systems. Journal of membrane science 130, 275-281.

Zydney, A. L. (2009). Membrane Technology for Purification of Therapeutic Proteins.

Biotechnology and Bioengineering 103, 227-230.

93

Appendix

Amino Acid Composition of Proteins

As discussed in Chapter 2, the protein surface charge density is determined by the

dissociation of the various ionizable amino acid residues on the surface of the protein.

The number and values of the various amino acids present in the proteins used in

this thesis are presented in Tables A.1 and A.2 below.

Table A.1 Number and values of charged amino acids in lysozyme (Sharma et

al., 2003)

Type

α-Amino 1 7.5

His 1 6.3

Arg 11 12.5

Lys 6 10.5

Glu 2 4.4

Asp 7 4

α-carboxyl 1 3.8

Tyr 3 9.6

94

Table A.2 Number and values of charged amino acids α-lactalbumin (Molek,

2008).

Type

N-term 1 9.87

His 3 6.04

Arg 0 12.5

Lys 12 10.54

Glu 8 3.9

Asp 9 3.9

C-term 1 2.16

Tyr 4 10.3

The small differences in values in Tables A.1 and A.2, e.g., 6.3 and 6.04 for

histidine, reflect the range of pKa values for the various amino acids reported in the

literature and likely reflect small differences in ionization potential associated with the

specific local environment within the different proteins. All calculations in this thesis

were performed using the respective values in Tables A.1 and A.2 since those numbers

have been shown previously to properly describe the net charge for lysozyme and α-

lactalbumin, respectively.