dynamic behavior of magnetic latex particles and

212
Louisiana State University LSU Digital Commons LSU Historical Dissertations and eses Graduate School 1994 Dynamic Behavior of Magnetic Latex Particles and Polyelectrolytes. Daewon Sohn Louisiana State University and Agricultural & Mechanical College Follow this and additional works at: hps://digitalcommons.lsu.edu/gradschool_disstheses is Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Historical Dissertations and eses by an authorized administrator of LSU Digital Commons. For more information, please contact [email protected]. Recommended Citation Sohn, Daewon, "Dynamic Behavior of Magnetic Latex Particles and Polyelectrolytes." (1994). LSU Historical Dissertations and eses. 5905. hps://digitalcommons.lsu.edu/gradschool_disstheses/5905

Upload: others

Post on 06-Jan-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Louisiana State UniversityLSU Digital Commons

LSU Historical Dissertations and Theses Graduate School

1994

Dynamic Behavior of Magnetic Latex Particles andPolyelectrolytes.Daewon SohnLouisiana State University and Agricultural & Mechanical College

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_disstheses

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion inLSU Historical Dissertations and Theses by an authorized administrator of LSU Digital Commons. For more information, please [email protected].

Recommended CitationSohn, Daewon, "Dynamic Behavior of Magnetic Latex Particles and Polyelectrolytes." (1994). LSU Historical Dissertations and Theses.5905.https://digitalcommons.lsu.edu/gradschool_disstheses/5905

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly from the original or copy submitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer.

H ie quality of this reproduction is dependent upon the quality of the copy submitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand comer and continuing from left to right in equal sections with small overlaps. Each original is also photographed in one exposure and is included in reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6” x 9" black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order.

A Bell & Howell Information Company 300 North Z eeb Road. Ann Arbor. Ml 48106-1346 USA

313/761-4700 800/521-0600

DYNAMIC BEHAVIOR OF MAGNETIC LATEX PARTICLES AND POLYELECTROLYTES

A Dissertation

Submitted to the Graduate Faculty o f the Louisiana State University and

Agricultural and Mechanical College in partial fulfillment o f the

requirements for the degree o f D octor o f Philosophy

in

The Department o f Chemistry

byDaewon Sohn

M .S., Hanyang University, Seoul, Korea, 1986 B.S., Hanyang University, Seoul, Korea, 1984

December 1994

UMI Number: 9524485

OMI Microform Edition 9524485 Copyright 1995, by UMI Company. All rights reserved.

This microform edition is protected against unauthorized copying under Title 17, United States Code.

UMI300 North Zeeb Road Ann Arbor, MI 48103

To

MiRang and DongWon

ii

ACKNOWLEDGMENTS

First I would like to especially extend my gratitude to Dr. Paul S. Russo, for his

invaluable support, thoughtfulness, guidance, and constructive criticism. Special

thanks are also extended to Drs. W. H. Daly, J. R. Collier, R. W. Hall, and A. R. P.

Rau for being my committee members. I sincerely appreciate their enthusiasm and

encouragement.

I wish to thank Dr. S. Lee, he advised me during my first year in Baton Rouge.

I am also grateful to Dr. Tahir Jamil and Russo's group members; Debbie, Zimei,

Keunok, Steve, Lucille, and Mike. Thanks are also given to Dr. Russo's family: Mrs.

Mary Russo, Michael, and Amy. I f they didn't invite me, I could have a couple o f

hungry Thanksgivings. Debbie, Steve, M ark, and Lucille, whom I consider the "Real

Macoy," helped me to correct the mistakes o f my English in this work. I would also

like to acknowledge Cindy Henk in the EM lab who took EM pictures, Dr. M. L.

Mclaughlin and Mr. A. Davila who labeled the polymer for the FPR experiment, Dr. D.

S. Poche' who measured the molecular weight o f NaPSS, and Dr. L. G. Butler and Mr.

Y. Lee who helped me to make the Helmholtz coil for the field induced study.

I am also indebted to Louisiana State University and NSF for their financial

support o f my graduate studies and researches.

To my parents, it would be hard to repay for your endless faith and support.

Likewise, I will always be indebted to my brother and sister. M y final sentence should

give to my wife, MiRang, who got married, had a baby, and prayed for me every single

day. For her incessant patience, love, and support I am forever thankful.

TABLE OF CONTENTS

DEDICATION.............................................................................................................................ii

ACKNOW LEDGM ENTS....................................................................................................... iii

LIST OF TA BLES................................................................................................................... vii

LIST OF FIG U R E S................................................................................................................viii

LIST OF SY M B O LS............................................................................................................. xiv

LIST OF A B BR EV IA TIO N S..............................................................................................xix

A B STR A C T............................................................................................................................. xxi

CHAPTER 1IN TRO D U CTIO N......................................................................................................................1

1.1 PROBE DIFFUSION STU DIES......................................................................... 21.1.1 GENERAL CONCEPTS............................................................................21.1.2 TECHNIQUE FOR PROBE DIFFUSION STUDIES........................ 4

1.2 CHARGED INTERACTIO N................................................................................51.2.1 COLLOID...................................................................................................... 51.2.2 POLYELECTROLY TES...........................................................................6

1.3 GENERAL CHARACTERISTICS OF M AGNETIC LATEX PA RTICLES............................................................................................................. 7

1.4 PROPERTIES OF MAGNETIC LATEX PA R TIC LE S................................91.4.1 SURFACE C H A R G E .................................................................................91.4.2 M AGNETIC SUSCEPTIBILITY.......................................................... 131.4.3 OPTICAL ANISOTROPY...................................................................... 14

1.5 REFEREN CES...................................................................................................... 17

CHAPTER 2PROBE DIFFUSION OF M AGNETIC LATEX PA RTICLES.................................... 20

2.1 BA CK G RO U N D .................................................................................................. 212.2 EX PERIM EN T......................................................................................................252.3 RESULTS AND DISCUSSION........................................................................ 26

2.3.1 IN DILUTE BINARY SO LU TIO N ..................................................... 262.3.2 IN PRESENCE OF POLY(STYRENESULFONATE)....................29

2.4 SU M M A R Y ...........................................................................................................312.5 REFEREN CES...................................................................................................... 37

CHAPTER 3INTERACTION BETW EEN POLYELECTROLYTES AND M AGNETICLATEX PA R TIC LE S............................................................................................................. 38

3.1 IN TR O D U CTIO N ............................................................................................... 39

3.1.1 COLLOID AND POLYM ER IN TER A C TIO N S.............................. 403.1.2 STATIC AND DYNAMIC LIGHT SCA TTERIN G ........................ 41

3.2 EX PERIM EN T...................................................................................................... 443.3 RESULTS AND DISCUSSIO N........................................................................ 48

3.3.1 INTERPARTICLE INTERACTIONS IN SALTCO N D ITIO N ............................................................................................48

3.3.1.1 PARTICLE CHARACTERIZATION IN PURE W ATER ... 483.3.1.2 INTERPARTICLE INTERACTIONS: NO ADDED

PO LYM ER.......................................................................................... 553.3.1.3 EFFECT OF PO LYELECTROLY TE......................................... 59

3.3.2 FLUORESCENCE PHOTOBLEACHING R E C O V ER Y ...............6 8

3.3.2.1 LABELING THE POLYSTYRENESULFONATE SODIUM SA L T .................................................................................6 8

3.3.2.2 CHARACTERIZATION OF LABELED PSS (L P S S ) 703.3.2 3 FLUORESCENCE PHOTOBLEACHING RECOVERY

M EASUREM ENT............................................................................. 723.3.2.4 STABILIZATION M EC H A N ISM .............................................. 73

3.4 SU M M A R Y ........................................................................................................... 763.5 R EFEREN CES.......................................................................................................79

CHAPTER 4KINETIC STUDIES OF M AGNETIC LATEX PARTICLES’ SELF ASSEM BLY UNDER APPLIED M AGNETIC FIELD ..................................................82

4.1 IN TR O D U CTIO N ................................................................................................ 834.2 KINETIC M ECHANISM AND THEORY..................................................... 84

4.2.1 M ECHANICAL C O N C EPTS................................................................ 844.2.2 PHASE TRANSITION C O N C E PT S...................................................854.2.3 KINETIC CONSIDERATIONS.............................................................8 6

4.3 VIDEO M IC R O SC O PY ......................................................................................8 8

4.3.1 GENERAL IN TR O D U CTIO N ..............................................................8 8

4.3.2 CELL GEOM ETRY AND M AGNETIC F IE L D ...............................8 8

4.3.3 SAMPLE PREPARATION.....................................................................904.3.4 IMAGE PR O C E SSIN G ...........................................................................914.3.5 RESULTS AND DISCUSSIO N.............................................................93

4.4 VIDEO SMALL ANGLE LIGHT SCATTERING..................................... 1004.4.1 GENERAL IN TR O D U CTIO N ............................................................1004.4.2 EXPERIM ENTAL SE T U P................................................................... 1014.4.3 DATA HANDLING AND EXPERIM ENTAL R E SU L T S 103

4.5 SU M M A R Y ......................................................................................................... 1104.6 REFEREN CES................................................................................................... 111

CHAPTER 5DYNAM IC BEHAVIOR OF A HIGH STRENGTH ROD-LIKEPOLYELECTROLYTE IN A STRONG ACID.............................................................. 114

5.1 IN TR O D U CTIO N ..............................................................................................115

v

5.2 EX PERIM ENTAL...............................................................................................1205.2.1 M A T E R IA L S...........................................................................................1205.2.2 M EA SU REM EN T.................................................................................. 121

5.3 R E S U L T S .............................................................................................................1225.3.1 SIGNAL CO N SID ERA TIO N .............................................................. 1225.3.2 SIMPLE AN A LY SIS..............................................................................128

5.3.2.1 POLARIZED M EA SU REM EN TS............................................ 1285.3.2.2 DEPOLARIZED M EA SU REM EN TS...................................... 1345.3.2.3 ANGULAR D EPEN D EN C E.......................................................134

5.3.3 CONTIN ANALYSIS W ITH M ULTIPLE R U N S........................ 1415.3.4 D ISCU SSIO N ...........................................................................................150

5.4 SU M M A R Y ......................................................................................................... 1535.5 R EFEREN CES.....................................................................................................154

CHAPTER 6

C O N C LU SIO N S.................................................................................................................... 156

APPENDIX A .......................................................................................................................... 159THE BEHAVIOR OF M AGNETIC LATEX PARTICLES IN CHEM ICALLYCROSS-LINKED G E L S....................................................................................................... 159

A .l IN TR O D U C TIO N ..............................................................................................160A. 2 EX PERIM EN TA L............................................................................................. 162

A.2.1 LATEX PA RTICLES............................................................................ 162A.2.2 SILICA G E L ............................................................................................ 163A.2.3 ACRYLAM IDE G EL............................................................................ 164A.2.4 M EA SU R EM EN T................................................................................. 164

A.3 RESULTS AND D ISC U SSIO N ..................................................................... 165A.3.1 ORDINARY LATEX PARTICLES IN SILICA G EL...................165A.3.2 M AGNETIC LATEX PARTICLES IN SILICA G E L ..................171A.3.3 M AGNETIC LATEX PARTICLES IN PAA G E L ...................... 171

A.3.3.1 TRANSLATIONAL D IFFU SIO N ............................................ 171A.3.3.2 ROTATIONAL D IFFU SIO N .....................................................176

A.4 SUM M ARY......................................................................................................... 176A. 5 R E FER E N C E S....................................................................................................179

APPENDIX B : UNITS FO R M AGNETIC PR O PER TIES.........................................181

APPENDIX C: COPYRIGHT PERM ISSIO N............................................................... 183

VITA 185

LIST OF TABLES

Table 1.1 Comparison between normal latex particle and magnetic latexparticle............................................................................................................................................ 9

Table 1.2 Description o f the magnetic latex particles which were used........................ 13

Table 3.1 Dynamic light scattering data o f three different magnetic latexparticles in pure water. Dynamic radii are from slope o f Vv (Rsw)> H v(RsHv), and intercept from H v (R1̂ ) . Unit: A...................................................................54

Table 3.2 Radius o f magnetic latex particles in NaPSS, salt solution.These values are calculated ffom intensity measurements, Vv and H vdynamic light scattering measurements. H sv, H!v mean slope and interceptfrom H v measurement, respectively....................................................................................... 58

Table 4.1 The relation between the magnetic strength (Gauss) and scalingparameters. Interaction parameter: X, sample response to the magneticfield: 1/t, and the scaling constan t....................................................................................... 99

Table 4.2 Comparison o f the rate o f the cluster size growth betweenSALS and microscopic observation. Unit: pm/s. It is hard to get the dataffom the conditions where the time scale is too fast (a) or too slow (b).....................109

Table 5.1 Samples o f PBO/M SA-M SAA with different concentrations and salt conditions. Intrinsic viscosity was measured with solvent viscosity,14.56 cP,without salt (NaMSA), at 25°C.......................................................................... 121

Table 5.2 Slow decay rate and corresponding apparent diffusion coefficient o f the low molecular weight (M = 19,000), at two different salt conditions, and 4 different concentrations......................................................................... 133

Table 5.3 The calculated translational and rotational diffusion ffom theKirkwood-Riseman approach. The unit o f translational diffusion is1 0 ' 8 cm2 s" 1 and that o f rotational diffusion is s" 1 ...........................................................151

Table 5.4 The dimension o f PBO ffom Odijk’s polyelectrolytesconsideration............................................................................................................................. 152

LIST OF FIGURES

Figure 1.1 Structure o f the magnetic latex particle. Filled circles representstyrene monomer, shadow circles represent sulfate radical, and emptycircles show the surfactant molecules................................................................................... 1 0

Figure 1.2(a) Transmission electron microscope picture o f the purified magnetic latex particles. Particles are polydisperse and (b) there are many dark spots. The light region in the center indicates that particles are preferentially located near........................................................................................................1 1

Figure 1.3 TG/DTA analysis o f the magnetic latex particles. The wt% o fthe magnet can be calculated by the weight loss o f the latex particles.......................... 1 2

Figure 1.4 The interaction parameter o f the particles which have charges,and magnetic interaction...........................................................................................................16

Figure 2.1 Light scattering alignment for the H v and Vv measurement....................... 22

Figure 2.2 Idealized model o f a magnetic latex particle (not to be confusedwith actual latex particles).......................................................................................................24

Figure 2.3 Decay rates ffom H v and Vv measurements vs. q2, for Seradynmagnetic latex in pure water. Apparent particle radii are indicated, alongwith linear correlation coefficients for the least squares lines......................................... 28

Figure 2.4 Decay rates ffom H v and Vv measurements vs. q2, forPolysciences magnetic latex in pure water. Apparent particle radii areindicated, along with linear correlation coefficients for the least squareslines...............................................................................................................................................30

Figure 2.5 Decay rates ffom Vv measurement vs. q2 for Polysciencesmagnetic latex particles in varying NaPSS concentration................................................ 32

Figure 2.6 Decay rates ffom H v measurement vs. q2 for Polysciencesmagnetic latex particles in varying NaPSS concentration................................................ 33

Figure 2.7 M acroscopic shear viscosity o f NaPSS solution, plus rotationaland translational micro-viscosity for Polysciences magnetic latex................................. 34

Figure 2.8 Intensity autocorrelation function (Vv) o f Seradyn magnetic latex in the presence o f an externally applied, oscillating magnetic field.The initial diffusion-like decay is followed by sinusoidal oscilations............................. 36

Figure 3.1 Magnetic latex particle concentration dependent decay rate vs.q2. Apparent translational diffusion coefficients did not change muchbetween lx lO '3% and lxlO"5% range o f concentrations..............................................47

Figure 3.2 Typical decay profiles for Bangs magnetic latex particles in w ater in V v and H v geometries. M easurements have been done at 60° scattering angle and at 25°C................................................................................................... 49

Figure 3.3(a) Seradyn magnetic latex particles' (1) decay rate vs. q2 plot o fVv and H v measurement. (2) T/q 2 vs. q2 plot......................................................................50

Figure 3.3(b) Polysciences magnetic latex particles' (1) decay rate vs. q2

plot o f Vv and H v measurement. (2) T/q2 vs. q2 p lo t........................................................ 51

Figure 3.3(c) Bangs magnetic latex particles' (1) decay rate vs. q2 plot o fVv and H v measurement. (2) T/q2 vs. q2 plot......................................................................52

Figure 3.4 Guinier plot o f the Bangs magnetic latex particles in pure water(square), MLP in the 0.1 M NaCl solution (circle), and M LP in the 0.1 MNaCl + polyelectrolytes solution (triangle), cNaPSS = 3 .29x l0 ‘3 g/ml............................57

Figure 3.5 Normalized autocorrelation function vs. lag time for Bangs magnetic latex particles/NaPSS/0.1 M NaCl ffom Vv and Hv geometries. M easurements have been done at 60° scattering angle and at 25°C..............................60

Figure 3.6(a) Vv and (b) H v measurements o f q-dependent decay rate o f magnetic latex particles in different NaPSS concentrations in pure w ater.....................61

Figure 3.7(a) Vv and (b) H v measurements o f q-dependent decay rate o fmagnetic latex particles in different NaPSS concentration in 0.10 M saltsolution........................................................................................................................................ 62

Figure 3.8 Vv measurement o f diffusion coefficient vs. NaPSSconcentration by varying NaCl condition.............................................................................63

Figure 3.9 H v measurement o f diffusion coefficient vs. NaPSSconcentration by varying NaCl condition.............................................................................64

Figure 3.10 Translational (left axis) and rotational diffusion (right axis) o fthe M LP varying salt concentration...................................................................................... 6 6

Figure 3.11. Mechanism o f the labeling the NaPSS with fluorescent dye.(a) First step is to substitute the sodium sulfonate with the sulfonyl chlorideand (b) then the sulfonylchloride is reacted with amine containingfluorophore ............................................................................................................................... 69

Figure 3.12 DLS comparison between LPSS and NaPSS in different saltconditions. At high salt condition, hydrodynamic radii o f LPSS and NaPSSare the same................................................................................................................................ 71

Figure3.13 FPR traces for LNaPSS (cLNaPSS = 1 .5x 10‘ 5 g/ml) in 1.8wt% M LP suspension. Top: exponential decay and single exponentialfitted curve. Inset: decay rate scales linearly with K2. Bottom: In(peakAC-baseline) vs. time p lot............................................................................................. 75

Figure 3.14 Self diffusion o f LNaPSS as a function o f the quotient o f MLPsurface area and the area required to bind all LNaPSS (see text). 0.2 pmM LP was used for this experiment except the last two (full) points forwhich 0.8 pm M LP was used..................................................................................................77

Figure 4.1 Cell geometry, Helmholtz coil, to induce the magnetic field...................... 89

Figure 4.2 Experimental set up for the video microscopy experiment......................... 92

Figure 4.3 M icroscopy observation for the evolution o f the magnetic latex particle (0.8 pm) under the magnetic field, 0.2%, 6 G, (a) 0 time, (b) after 5 s,.(c) after 30 s, (d) after 50 s..............................................................................................94

Figure 4.4 Cluster size growth for 0.2 wt% o f MLP under differentmagnetic fields............................................................................................................................96

Figure 4.5 Cluster size growth for 0.1 wt% o f MLP under differentmagnetic fields............................................................................................................................97

Figure 4.6 Pow er law grow th for (a) 0.2 w t% (b) 0.1 w t% o f MLP under different magnetic fields, x is a constant for each different field....................................98

Figure 4.7 Video small angle light scattering setup........................................................ 102

Figure 4.8 VSALS observation for the evolution o f the magnetic latexparticle (0.8 pm) under the magnetic field......................................................................... 104

Figure 4.9 Raw intensity profile o f SALS o f MLP under the magneticfield, 0.5 A/ 6 G, 0.05 wt% , 670 nm.................................................................................... 106

Figure 4.10 Raw intensity profile o f SALS o f MLP under the magneticfield, 0.5 A / 6 G, 0.2 wt% , 670 nm...................................................................................... 107

Figure 4.11 ln(I/Imax) vs. ln(q/qmax) plot. Background intensitysubtracted, length scale = 27t/q.............................................................................................108

Figure 5.1 Chemical structure o f cis-PBO. At least tw o sites o f N or Oare protonated by strong acids..............................................................................................117

x

Figure 5.2 Intrinsic viscosity o f PBO/M SA-M SAA as a function o f salt.D ata file is provided by Dr. D. B. Roitman at Dow's Walnut Creek group in California...............................................................................................................................118

Figure 5.3 Data handling with single short run and multiple runs. Single run has high frequency noise at short lag time, but multiple runs have quiet enough to measure a very rapid decay. Data was analyzed by the in house developed program ALVAN................................................................................................ 123

Figure 5.4(a) Correlation function for PBO/M SA-M SAA solution: M =19,000; Cone. = 4 x l0 ‘ 3 g/ml; cs = 0.4 M; Vv measurement; 90° scatteringangle; 488 nm wavelength; 800/200 pin-hole setting; 300 s acquisition timesingle run; 25°C. (b) corresponding semilog representation.........................................125

Figure 5.4(c) Correlation function with strongly scattered latex, 0.087 pm, solution under the same pin-hole setting (800/200), X0 = 488 nm. (d) corresponding semilog representation................................................................................ 126

Figure 5.5(a) Correlation function for low salt condition o f PBO/MSA- M SAA solution: M = 19,000; Cone. = 4 x l0 " 3 g/ml; cs = 0 M; Vv measurement; 90° scattering angle; 488 nm wavelength; 800/200 pin-hole settings; 1,000 s acquisition time single run; 25°C....... 127

Figure 5.6(a) Correlation functions for high molecular weight PBO/MSA- M SAA solution and their salt dependence: M = 78,000; Cone. = 4x10" 3

g/ml; Vv measurement; 90° scattering angle; 632.8 nm wavelength;1,000/100 pin-hole settings; 1500 s acquisition............................................................... 129

Figure 5.6(b) Corresponding semilog plot o f Figure 5.6(a) for 0.45 M salt solution, and (c) for 0 M salt solution................................................................................ 130

Figure 5.7(a) Autocorrelation functions for low molecular weight PBO/M SA-M SAA solution and their salt dependence: M = 19,000; Cone.= 4x10 ‘ 3 g/ml; Vv measurement; 90° scattering angle; 488 nm wavelength;800/200 pin-hole settings; 300 - 1,000 s ........................................................................... 131

Figure 5.7(b) Decay rate distributions corresponding to Figure 5.7(a), and amplitude x f(A ) 1 / 2 vs. decay rate plot............................................................................... 132

Figure 5.8(a) Correlation function o f depolarized measurement o f PBO/M SA-M SAA solution and their salt dependence: M = 19,000; Cone.= 4x10 ' 3 g/ml; H v measurement; 90° scattering angle; 488 nm wavelength;800/200 pin-hole settings; 1200 s acquisition tim e ......................................................... 135

Figure 5.8(b) Corresponding decay rate o f Figure 5.8(a) 136

Figure 5.9(a) Angular dependence o f the autocorrelation functions for the high molecular weight PBO/M SA-M SAA solution at high salt condition:M = 78,000; Cone. = 4 x l0 ‘ 3 g/ml; cs = 0.45 M; Vv measurement; 6

different angles; 632.8 nm w avelength.................................... 138

Figure 5.9(c) Angular dependence o f the autocorrelation functions for the high molecular weight PBO/M SA-M SAA solution at low salt condition:M = 78,000; Cone. = 4 x l0 ' 3 g/ml; cs = 0 M ; Vv measurement; 6 differentangles; 632.8 nm wavelength; 1 ,000/100........................................................................ 139

Figure 5.10 T vs. q2 plot (bottom) and T/q 2 vs. q2 plot (top) o f the thirdcumulant o f Vv autocorrelation functions. High molecular weight(78,000), high concentration 4x10 ' 3 g/ml and different salt conditions(0.01, 0.10, 0.45 M ), at 25°C............................................................................................... 140

Figure 5 .11(a) PRO CONTIN analysis o f low molecular weight PBO/M SA-M SAA solution at high salt condition: M = 19,000; Cone. =4 x l0 ' 3 g/ml; cs = 0.45 M; Vv measurement; 50° scattering angle; 514.5 nm wavelength; 600/100 pin-hole setting; 15,000 s...............................................................142

Figure 5 .11(b) PRO CONTIN analysis o f low molecular weight PBO/M SA-M SAA solution at high salt condition: M = 19,000; Cone. =4 x l0 ' 3 g/ml; cs = 0.45 M; Vv measurement; 60° scattering angle; 514.5 nm wavelength; 600/100 pin-hole setting; 15,000 s............................................................... 143

Figure 5 .11(c) PRO CONTIN analysis o f low molecular weight PBO/M SA-M SAA solution at low salt condition: M = 19,000; Cone. =4 x l0 ' 3 g/ml; cs = 0 M ; Vv measurement; 60° scattering angle; 514.5 nm wavelength; 800/200 pin-hole setting; 15,000 s acquisition..........................................144

Figure 5 .11(d) PRO CONTIN analysis o f depolarized scattering for low molecular weight PBO/M SA-M SAA solution at high salt condition: M =19,000; Cone. = 4 x l0 " 3 g/ml; cs = 0 M; H v measurement; 60° scatteringangle; 514.5 nm wavelength; 800/200 pin-hole............................................................... 145

Figure 5 .11(e) EXSAM P analysis o f low molecular weight PBO/MSA- M SAA solution at high salt condition: M = 19,000; Cone. = 4x10 ' 3 g/ml; cs = 0.45 M; Vv measurement; 60° scattering angle; 514.5 nm wavelength;600/100 pin-hole setting; 15,000 s acquisition................................................................ 146

Figure 5.12 Angular dependence o f the slow decay mode o f low molecular weight PBO/M SA-M SAA at high salt condition. M = 19,000; Cone. =4x10 ‘ 3 g/ml; cs = 0.45 M; Vv measurement; 6 different scattering angles;514.5 nm wavelength; 600/100 pin-hole settings.......................................................... 148

xii

Figure 5.13 Angular dependence o f the fast decay mode for low molecular weight PBO/MSA-MSAA. M = 19,000; Cone. = 4 x l0 ' 3 g/ml; cs = 0.45 M and 0 M; Vv and H v measurement; 4-5 different scattering angles; 514.5 nm wavelength; 600/100 pin-hole se ttings ....................................................................... 149

Figure A. 1 Diffusion coefficient change o f the silica sol-gel transition with time. Sample composition was 0.5 ml TEOS, 10 ml EtOH, catalyzed with NH 4 OH. Light scattering measurement was done at 25°C, 90° scattering angle............................................................................................................................................166

Figure A.2 Correlation function o f polystyrene latex particles, diameter =0.059 pm, in the silica gel network, compared to that in the water. Thecomposition o f the gel is 0.5 ml TEOS, 10 ml EtOH, catalyzed withNH 4 OH. M easurement was performed at 2 5 °C .............................................................168

Figure A.3 Correlation function o f the polystyrene latex particles, diameter = 0.059 pm, in different TEO S/EtO H ratio o f completed silica gel.TEO S/EtO H ratio are 0.5%, 10%, 12%, 15%, respectively......................................... 169

Figure A.4 Diffusion coefficient o f latex particles in silica gel decreasing by increasing the TEO S/EtO H molar ratio. D° = 7 .74xl0 ‘ 8 cm2/s ............................. 170

Figure A. 5 Autocorrelation function o f the magnetic latex particles during the silica sol-gel transition. There was no correlation function after 24 hours, which means the M LP are trapped in the gel. The gel condition was 0.5% T E O S ..............................................................................................................................172

Figure A . 6 The time averaged autocorrelation function from the 50 different spots o f the M LP/PAA gel system. The particles' autocorrelation function is coming from the polarized light scattering in 3 different concentration (2.5, 3.5, 4.5 w t% ) o f P A A ......................................................................174

Figure A. 7 The ensemble averaged autocorrelation function from the 50 different spots o f the M LP/PAA gel system. The particles' autocorrelation function is coming from the polarized light scattering in 3 different concentration (2.5, 3.5, 4.5 w t% )......................................................................................175

Figure A . 8 The depolarized autocorrelation function from the 50 different spots o f the M LP/PAA gel system. The particles' autocorrelation function is coming from the depolarized light scattering in 3 different concentration (2.5, 3.5, 4.5 wt% ) o f P A A .................................................................................................. 177

LIST OF SYMBOLS

a spacing between the particles

b distance between charges along the chain

B baseline o f the autocorrelation function

B magnetization

Be Bjerrum length

c D viscous drag coefficient

c concentration o f solution

Cs concentration o f salt

D app apparent diffusion coefficient

Db diffusion coefficients o f LNaPSS that are bound

Dc charge density

D° diffusion in pure solvent

D°t translational diffusion in the simple solvent

Df diffusion coefficients o f LNaPSS that are free

Dm mutual diffusion coefficient

Dp probe diffusion coefficient

Dr rotational diffusion coefficient

D s self diffusion coefficient

D s diffusion coefficient for clusters o f size s

D t translational diffusion coefficient

d particle diameter

d fringe spacing in forced Rayleigh scattering

deff effective electrostatic diameter surrounding the chain

df fractal dimension

xiv

dH electrostatic diameter o f charged rod-like polyelectrolytes

d 0 diameter o f uncharged rod

dt spacing between the tubing o f Helmholtz coil geometry

e charge o f an electron

Fm magnetic force

Fv viscous drag

f friction coefficient

fm mutual friction coefficient

ft friction factor associated with translational motion

f(A) instrumental param eter in autocorrelation function

f^A)max maximum value o f f(A)

f(q,x) dynamic structure factor

C^2\x ) intensity autocorrelation function

g(2 )(x) normalized autocorrelation function

g^)(x) normalized field autocorrelation function

H applied magnetic field

Ha Hamaker constant

H v detection o f horizontally polarized component o f scattered light withvertically polarized incident beam

I intensity o f scattered light

I ionic strength

/ electric current

IHv intensity o f H v measurement

IVv intensity o f V v measurement

K spatial frequency in fluorescence photobleaching recovery

Kp Porod constant

xv

K' bulk modulus

k Boltzmann constant

L striped pattern spacing in fluorescence photobleaching recovery

L contour length o f the polymer

M magnetization

M molecular weight

N number o f discrete time intervals in autocorrelation function

N number o f ions per cm 3

N a Avogadro's number

N (t) number o f clusters on a certain time scale

n solution refractive index

P parking area

Ps polymer bulk density

Q persistence length o f the polyelectrolytes

q scattering vector

R radius o f the particle

Rb radius o f gyration

R l Hv radius which is from intercept o fI7 q 2 plot with H v geometry

R s h v radius which is from slope o f r/q2 plot with H v geom etiy

>>P4 radius which is from slope o f r/q2 plot with Vv geometry

h the radius o f the tubing o f Helmholtz coil geometry

S surface charge density

s cluster sizes on a certain time scale

s(t) cluster size distribution function

T temperature

% aggregation time (scale)

xvi

coarsening time (scale)

potential energy

drift velocity

attractive energy

repulsive energy

vertically polarized component o f the scattered light for a vertically polarized incident beam

mole fractions o f LNaPSS that are bound

mole fractions o f LNaPSS that are free

exponent constant

ion valency

optical anisotropy

magnetic susceptibility

dielectric constant o f the solvent

van der W aals interaction

volume fraction o f the solute

polymer volume fraction

effective solid surface fraction

decay rate

rate at which the wave amplitude o f the fundamental component decays

decay rate o f rotational motion

decay rate o f translational motion

decay rate from H v geometry

decay rate from Vv geometry

Debye screening length

viscosity o f the solution

translational microviscosity

rotational microviscosity

magnetic interaction parameter

dipole-dipole interaction

laser wavelength in vacuo

dipole moment

shear modulus

magnetic permeability o f the vacuum

osmotic pressure

angle between the incident and scattered light

time delay (lag time) in dynamic light scattering

surface potential

linear charge density

correlation length, mesh size in the gel

zeta potential

number density

critical concentration where the rods start to overlap

osmotic compressibility

xviii

LIST OF ABBREVIATIONS

AA acrylamide, CH2 =CHCONH 2

AC(t) magnitude o f the ac signal

ALV CONTIN CONTIN analysis with ALV correlator

AP ammonium persulfate, (NH 4 )2 S2 0 8

BAA N, N,-methylene bisacrylamide, (H 2 C=CHCONH)2 CH 2

CC theory Manning's counterion condensation theory

CONTIN Laplace inversion algorithms

DLA diffusion limited aggregation

DLCA diffusion limited cluster aggregation

DLS dynamic light scattering

DLVO Deijaguin, Landau, Verwey & Overbeek theory (originators o f thetheory o f colloidal stability)

ER electrorheological (fluid)

EXSAM P Laplace inversion algorithms with first order smoothing

FPR fluorescence photobleaching recovery

FRS forced Rayleigh scattering

HPC hydroxypropyl cellulose

K R Kirkwood-Riseman (approach for transport)

LP latex particle

LS light scattering

MLP magnetic latex particle

MOS metal oxide semiconductor

M SA methanesulfonic acid, CH 3 S 0 3H

M SAA methanesulfonic acid anhydride, (CH 3 S 0 2)20

NaPSS sodium poly(styrene sulfonate)

NM R nuclear magnetic resonance

PB theory Poisson-Boltzman theory

PBO poly(p-phenylene cis-benzobisoxazole)

PBT poly(p-phenylene trans-benzobisthiazole)

PEO poly(ethyleneoxide)

PFA tetrafluoroethylene-perfluoroalkylvinylether copolymer

PM M A poly(methylmethacrylate)

PRO CONTIN CONTIN analysis originated from Provencher

PTFE poly(tetrafluoroethylene)

QELS quasi elastic light scattering

RLA reaction limited aggregation

RLCA reaction limited cluster aggregation

SALS small angle light scattering

SAXS small angle X-ray scattering

SIT silicon intensified target

SLS static light scattering

TEM transmission electron microscope

TEM ED N, N, N', N'-tetramethylethylenediamine, (C H ^^N C H ^C F^N C C H ^

TEOS tetraethoxysilane, Si(OC2 H 5 ) 4

TG/DTA thermal gravimetry/differential thermal analysis

THF tetrahydrofuran, C4 H80

VSALS video small angle light scattering

ZADS zero-angle depolarized light scattering

XX

ABSTRACT

The dynamic behavior o f magnetic latex particles has been explored. Polarized

and depolarized dynamic light scattering from binary and ternary systems o f these

particles have provided rotational and translational difliision coefficients. M icroscopic

views o f rotational and translational diffusion o f these particles are compared with

macroscopic viscosity from the Stokes-Einstein equation (Chapter 2).

The polymeric stabilization o f magnetic latex particles has also been

investigated by static and dynamic light scattering. Using the optical anisotropy o f

magnetic latex particles, the translational and rotational diffusion coefficients o f the

particles under various salt conditions were determined. The stability o f the

superparamagnetic latex particles depends on electrostatic repulsion and van der Waals

attraction. Translational and rotational diffusion o f the magnetic latex particles

decrease abruptly in the high salt condition, but are recovered upon addition o f a

polyelectrolyte polymer. Polystyrene sulfonate sodium salt stabilizes the flocculated

particles but restricts their motion. Self diffusion studies with fluorescence

photobleaching recovery have been done with labeled NaPSS to verify that the stability

arises from a mechanism other than conventional steric stabilization (Chapter 3).

Applied magnetic fields induce the end-to-end attachment o f the magnetic latex

particles. Kinetic growth o f these particles under a magnetic field has been studied by

optical microscopy and small angle light scattering. Average cluster sizes determined

from the microscopy images and the SALS patterns have been compared (Chapter 4).

The polyelectrolyte studies were extended to a high strength rod-like

polyelectrolyte system (PBO/M SA-M SAA). Slow polymer chain diffusion and very

rapidly decaying intensity autocorrelation functions were measured by depolarized and

polarized light scattering data (Chapter 5).

xxi

The dynamics o f trapped magnetic latex particles in porous silica gel and

acrylamide gel depend on the gel structure and its viscoelastic properties. The

translational and rotational diffusion o f the magnetic latex particles and the ordinary

latex particles inside the gel netw ork have been investigated preliminarily with dynamic

light scattering (Appendix A).

CHAPTER 1

INTRODUCTION

1

2

1.1 PROBE DIFFUSION STUDIES

1.1.1 GENERAL CONCEPTS

Probe diflusion means tracing probe molecules, normally in ternary (probe:

polymer matrix: solvent) systems. These studies produce much information about

configurational and conformational structures o f the diffusing polymers, their dynamics,

and the structure or morphology o f the matrix polymers. From the experimental point

o f view, probe molecules should have dominant scattered intensity compared to the

matrix polymer, or selective optical signal compared to the background molecules.

M any different probes and matrix polymers have been used. Polymer latex particles

[1.1], proteins [1.2], poly(tetrafluoroethylene) copolymer [1.3], and labeled molecules

[1.4] are reported as probes. Hydroxypropyl cellulose [1.5], poly(methylmethacrylate)

[1.6], poly(ethyleneoxide) [1.7], bovine serum albumin [1.8], dextran [1.9], and even

acrylamide gel are used as matrix polymers.

The probe diffusion coefficient, Dp> is a dynamic parameter o f probe molecules

in a solvent or through a matrix. There are tw o different diffusion coefficients. One,

which is called the mutual or cooperative diffusion coefficient, Dm, is based on the

concentration gradient and is usually measured by dynamic light scattering (DLS). The

self diffusion coefficient, Ds, reflects the random diffusion o f individual molecules in the

absence o f any gradient. Optical tracer methods such as forced Rayleigh scattering

(FRS) [ 1.10] or fluorescence photobleaching recovery (FPR) [ 1.11 ] are used for these

studies. Pulsed-field-gradient NM R [1.12] is another technique to determine the self

diffusion o f the molecule.

The mutual diffusion coefficient, Dm, is the hydrodynamic and thermodynamic

response o f a collection o f macromolecules to a concentration gradient, and is given by

( 1.1)

The osmotic compressibility, (—) , is the driving force for the polymer motion, fm is' dc* p j

the mutual friction coefficient, and <j> is the volume fraction o f the solute.

Hydrodynamic friction opposes the thermodynamic concentration gradient. At low

concentration this equation simplifies to the Einstein form D° = kT/f. Many

experimental results show the scaling relations between the probe diffusion coefficient,

Dm or Ds, and the diffusion coefficient at c = 0, D°. Some important results are

summarized in the following.

Mutual diffusion in binary solution f 1.131

D m = D 0(l + a J ) ) (1.2)

C V * 1.45

Self diffusion in binary solution \ 1 .141

D s = D0(l + a s<j>) (1.3)

a s « -1 .8 3

Probe diffusion in ternary system [1 .151

D p = D 0 e x p ( - a c v) (1.4)

v = 0.5 to 1.0

a ^ M 1, R 8 y: 0 .8-1.0,5: 0.0+0.2 (for large probes)

where o ^ , a s, v, y, and 5 are a constant, M is the molecular weight, and R is the radius

o f the polymer.

When matrix systems consist o f polyelectrolytes or polymer gel networks, the

probe diffusion studies are more complicated, and one should consider the stabilization

o f the probe molecules, charge interaction, viscoelastic modulus o f the gel network,

etc. A preliminary study o f particle motion in a gel [1.16] showed the following scaling

equation,

D p » e x p (-a R s) (1.5)

5 = 0.5 - 2 .0

1.1.2 TECHNIQUE FOR PROBE DIFFUSION STUDIES

M utual diffusion is commonly measured by dynamic light scattering. Brownian

motions o f polymer molecules cause phase and polarization changes o f scattered light,

and the scattered intensity will fluctuate with time, typically on a ps to ms time scale.

This intensity fluctuation gives information on the dynamic motions o f the molecules.

The intensity autocorrelation function, which describes how an intensity fluctuation

relaxes back to average intensity, is

G (2 )(x) = ( l( t) I ( t + x)}

= ^ : Z I ( t i)I(ti + x ) ( 1 .6 )N i=l

where N is the number o f samples (number o f discrete time intervals) and t is the time

delay or "lag time." At zero time delay, G ^ (x ) is the mean square intensity <I2 (t)>,

and at large time delay

lim ( l( t) I ( t + x)) = ( l ( t ) ) 2 (1.7)I - > o O

G ^ (x ) is maximal at x = 0 and decays toward large x. The normalized form o f the

Equation 1.6 is

The decay o f the autocorrelation function is related to the diffusion coefficient; further

relationships will be developed in Chapter 2.

Typical techniques to measure self diffusion are forced Rayleigh scattering

(FRS) and fluorescence photobleaching recovery (FPR). Both techniques usually need

labeled molecules such as photochromic or photobleachable polymers to trace the

diffusion o f the polymers. These labeled polymers are photoexcited by illuminating

samples with fringe patterns. In case o f FRS [1.17], these fringe patterns are made by

5

the interference o f tw o laser beams, which is produced by splitting the laser beam,

recombining, and varying the angle between the tw o beams, and by a Ronchi ruling

(coarse diffraction grating) in FPR. The periodic distribution o f bleached molecules

decays with time, and the relaxation time is given by

x = - ^ ~ : FRS (1.9a)4tc2 Ds

d. fringe spacing,

D s: self diffusion

IJx = —-5— : FPR (1.9b)

4ft DsL: striped pattern spacing

Other methods have been introduced for the self-diffusion studies such as spin

echo N M R [1.18], and electric field light scattering [1.19].

1.2 CHARGED INTERACTION

1.2.1 COLLOID

The stability o f the charged particles was first studied by two different groups in

1950's, Deijaguin and Landau o f the Soviet Union, and Verwey and Overbeek o f the

Netherlands. Their combined theory suggested that the stability o f colloid particles

depends on the balance between van der W aals attraction (VA) and electric repulsion

energies (VR).

The van der Waals repulsion between the colloidal particles is similar to a

London treatment o f the atoms. It was developed by Ham aker and the attraction

energy is given by

v * = i S ? ( U 0 )

where is the effective Hamaker constant and h is the distance between the particles.

The magnitude o f VA is o f order kT for the latex particles.

6

The electric repulsion energy between the charged sphere with radius R is

where e is the dielectric constant o f the solvent, vj/ 0 is the surface potential which is

normally substituted by the £ potential in aqueous systems. The actual potential

measured by electrophoretic mobility, and is proportional to the C, potential, C, =

(4 7 it i / e ) u where e is the dielectric constant and u is electrophoretic mobility (~ 10- 4

cm 2 /v.s) [1.20], The thickness o f the double layer, or Debye length, is

where N is number o f ions per cm3, z is the ion valence, and e is the electrostatic

charge.

1.2.2 POLYELECTROLYTES

The theory o f colloid stability is based on the sphere or plate model. However,

the charged interactions among real polymers are more complicated. One m odem

theory o f polyelectrolyte solutions is based on the model o f an infinitely long cylinder

with electric charge distributed uniformly along its surface (polyion). The solution also

contains salts such as sodium and magnesium which can neutralize the polyion by

condensation onto its surface. It also has uncondensed counterions as well as co-ions.

For this system, the counterion condensation (CC) theory is developed as a simple,

analytical alternative to the Poisson-Boltzman (PB) theory. Manning [1.21] proposed

that condensation on the polyion occurs until the net linear charge density o f the

polyion is reduced below a critical value. In this case the key parameter is the linear

charge density (£,), which is

( 1.12)

ekTb( 1.13)

7

where e is the electric charge, e the bulk dielectric constant, and b is the average

spacing between charges. The theory states that if E, is less than unity, the interaction

o f small ions with the polyions may be treated in the Debye-Hlickel approximation. I f E,

exceeds unity, sufficiently many counter-ions will condense on the polyion to lower the

net value o f the charge-density parameter E, to unity. The electrostatic component to

the persistence length o f the polyelectrolyte based on Manning's CC theory is given by

( U 4 )

where A is the segment length o f the polymer. This is the case o f A < Be, where Be is

the minimum distance between the condensed charges, called the Bjerrum length, wliich

is obtained by setting E, = 1 in Equation 1.13 and solving for b, i. e. B e = e2 /ekT. The

general relation for the persistence length o f polyelectrolytes given by Odijk and

Fixman [1.22], is

Q = W < U 5 )e2

where Be = ------ . The dimension o f the charged polymer is calculated by theseekT

equations. The electrostatic diameter o f a charged rod-like polyelectrolyte is [1.23]

d H = d 0 + K _1 In/ 47rB.

* ' + 0.077b k

(1.16)

where d0 is the diameter o f the uncharged rod.

1.3 GENERAL CHARACTERISTICS OF MAGNETIC LATEX PARTICLES

Latexes are aqueous suspensions o f polymer particles prepared by emulsion

polymerization [1.24], Latex particles are grown by the following steps: Micelle

Formation Step, where hydrophobic and hydrophilic end groups o f emulsifier

molecules make the micelle in the aqueous phase; Swollen Micelle Structure Step,

8

where the monomer (e.g. styrene) is dispersed and dissolved within the micelle

structure; Polymer Precursor Step, where a free radical initiator makes the polymer

precursor (oligomers); Polymer Latex Step, where monomers are polymerized and

form a swollen micelle (polymer particles). The size o f these particles is 0.05 ~ 10 pm.

The extraordinarily uniform shape allows them to be used as calibrated particles, and

they have well known industrial applications such as binders in paint, adhesives, paper

coatings, etc. The surfaces o f latex particles can be covered with various chemical

groups, mainly with sulfonate and carboxylate moieties. Particle interactions depend

on van der Waals attraction and electric charged repulsion.

M agnetic m icrospheres are composite particles made o f polystyrene and finely

divided magnetic iron oxide. Ideally, the magnetic pigment is evenly distributed

throughout each particle as fine crystals o f 1 0 - 2 0 0 A in diameter. M agnetic iron oxide,

which is organically dispersible, is finely dispersed throughout a monomer phase with a

polymerization initiator added [1.25]. Figure 1.1 shows the structure o f the magnetic

latex particles, MLP. Even though the shape o f the MLP are geometrically isotropic,

commercialized magnetic particles are not perfectly monodispersed (Figure 1.2).

M agnetite crystals imbedded in the latex structure are randomly distributed, but the

randomness o f the big particles is greater than that o f small particles. The magnetic

content o f the particles is determined by thermal gravimetry. After passing the

decomposition point o f polystyrene latex, only solid magnets remain on the TG/DTA

chamber, as shown in Figure 1.3. These particles exhibit superparamagnetic properties;

that is, magnetization, B, o f the particle increases with the applied magnetic field, H,

but falls back to zero if the field is removed. W hen the field is removed, the particles

demagnetize and redisperse perfectly. The superparamagnetic property o f the particles

is often used in biomedical separations. Proteins can be fixed on the surface either by

passive adsorption or, preferably, by covalent linkage. Then protein attached particles

9

are attracted by the magnetic field. Due to relative high density (1.2 -2.2 g/ml,

compared to ~ 1 . 1 g/ml for normal latex) o f the magnetic latex, the particle settles

rather quickly. The normal sedimentation velocity for 1 pm M LP is around 0.5 pm/s.

Magnetic latexes which w ere used in this experiment contain the surfactant (sodium

dodecyl sulfate at 5 g/1) to impart long-term storage stability.

Table 1.1 Comparison between normal latex particle and magnetic latex particle.

Size

(pm)

Density

(g/ml)

Surface

group

Color Uniformity Use

Latex 0 .0 2 - 1 0 0 1.05-1.5 - n h 2, - c o o h White MonodisperseCalibration

Standard

MagneticLatex

0.05-2.6 1 .5-2.0 - n h 2, - c o o h Brown Polydisperse Biomedical

Separation

1.4 PROPERTIES OF MAGNETIC LATEX PARTICLES

1.4.1 SURFACE CHARGE

The charge density o f each particle and the parking area, which is defined as the

micro particle surface area occupied by a single functional group, are useful to

determine the charged interactions o f the particle. The number o f ionized groups on

the surface can be determined by conductometric titration. Sodium hydroxide titrant is

added until an equivalence point is reached. The surface charge density and parking

area are defined as follows [1.26],

S = (1.004)DcdPs (1.17)

P = l / S (1.18)

10

Surfactant

m Sulfate Radical

F e30 4 Iron Oxide

Styrene

Figure 1.1 Structure o f the magnetic latex particle. Filled circles represent styrene monomer, shadow circles represent sulfate radical, and empty circles show the surfactant molecules.

(b)0 .2 pm

Figure 1.2(a) Transmission electron microscope picture o f the purified magnetic latex particles. Particles are polydisperse and (b) there are many dark spots. The light region in the center indicates that particles are preferentially located near the surface.

12

600

156.2 C -34 ug

500

400

-10W eight Changeuo

c i 300£<D

H

200-15

-20100

Temp.

576.2 C -2320 ug

-25

6020 400

Time/min.

Figure 1.3 TG/DTA analysis o f the magnetic latex particles. The wt% o f the magnet can be calculated by the weight loss o f the latex particles.

Sh £1

/01

13

where S is the surface charge density (charge group/A2), Dc is the charge density

(meq/gr), Ps is the polymer bulk density, d is the particle diameter, and P is the parking

area. A parking area o f 200 indicates that for every 200 A2, there is just one COOH

group. Samples which were used in this experiment are listed in Table 1.2.

Table 1.2 Description o f the magnetic latex particles which were used.

SampleName

Diameter(pm)

Mag % Density(g/ml)

SurfaceSurfaceCharge(lieq/g)

Parking Area

(sq A/grp)

LotNumber

Bangs 68 0.8 68 2.27 -COOH 197 2.7 M l-070/60/380

Bangs 42 0.8 42 1.57 -COOH 97 8.1 M l-070/40/385

Seradyn 0.7 58 - -S 03 - -M l-

070/60/319

Bangs 005 0.05 95 4.28 -COOH - -M0000501

CN

Polvsciences 0.05 12 -SO, . . 404206

1.4.2 MAGNETIC SUSCEPTIBILITY

Magnetizable substances can be grouped into several types depending on the

way they are magnetized by the external field. Each o f following quotation is from ref

[1.27], Diamagnetic substances "acquire a magnetization opposed to the magnetic

field and have a magnetic permeability less than that o f a vacuum." Paramagnetic

substances acquire "a magnetization in the same direction as the magnetic field, which

has small but positive susceptibility varying but little with magnetizing force (such as

aluminum)." Superparamagnetism is analogous to paramagnetism except that the

magnetic moment is much larger; the existence o f superparamagnetic materials was first

predicted by Neel [1.28]. Ferromagnets exhibit "a permanent magnetization even in

14

the absence o f an external magnetizing force." Antiferromagnets (such as manganese

monoxide) "give zero net magnetization because o f antielectronic spins and will not be

oriented by an external magnetizing field." Ferrimagnets give "net magnetization due

to atomic or ionic magnetic moments in the same direction," and acquire a resultant

moment in one direction, and a weak ferromagnetism appears [1.29].

The magnetizable particle will acquire a dipole proportional to the external field

H,

p = p 0 (^)7rR \ H (1.19)

where R is the sphere radius, % is magnetic susceptibility, and p 0 is magnetic

permeability o f the vacuum. Under low enough magnetic field, 104 A/m or less, their

magnetization is proportional to the external field M = %H. M agnetic susceptibility

describes the response o f a particle to an external magnetic field. It can be obtained by

a magnetophoresis experiment, "balancing the magnetic force Fm = p 0 |aV // with the

viscous drag

Fv = r|uC D, where ri is the viscosity o f the suspending liquid, u is the measured drift

velocity, and CD is a drag coefficient depending on the size and shape o f the particles.

[1.30]" The magnetic susceptibility o f the Bangs 6 8 sample which is the most heavily

used particle in this study is 1.66xl0 ' 2 emu/Oe cm 3

Interactions between particles are governed by magnetic interaction, van der

Waals attraction, electrostatic repulsion, and weak gravitational attraction. Figure 1.4

shows the potential energy versus surface-to-surface separation o f sterically protected

colloidal particles.

I.4.3 OPTICAL ANISOTROPY

Uniform latex particles are optically isotropic. However, magnetic latex

particles have a partial crystalline structure which changes the intrinsic polarizability o f

the particle and makes them optically anisotropic. The anisotropy o f the structural

15

units rotates the induced dipole moment away from the incident polarization direction

and leads to a depolarized light scattering signal. The depolarized scattered light from

the anisotropic properties is used to determine rotational motion o f the particles.

The time correlation function o f the depolarized intensity which has optical

anisotropy, p, is

G(2 )(x) = A + B|g(1)(x ) | 2 = A + B e x p [-2 (q 2 Dt + 6 D r)x)] (1.13)

where A is a constant, the constant B is proportional to P2, and Dt and Dr are the

translational and rotational contribution, respectively. For particles in the size range

considered here, the decay rate associated with rotational diffusion is normally faster

than that associated with translational diffusion, so experiments are sometimes difficult

to perform. Zero angle depolarized scattering experiment in which there is no

dependence on the translational diffusion (q2 Dt = 0), is sometimes performed to study

rotational motion [1.31]. Fabry-Perot interferometry is another choice for the fast

depolarized experiment.

Piazza et al. reported anisotropic properties o f the latex particle PFA

(tetrafluoroethylene-perfluoroalkylvinylether copolymer) which has a partially

crystalline internal structure [1.32]. His particles have electro-optical and non-linear

optical properties due to their crystallinity and optical anisotropy. Our lab has

performed the depolarized experiment with "Fluon" PTFE particles, and with magnetic

latex particles [1.33], The optical anisotropy o f the M LP is caused by the embedded

magnet crystallites. The focus o f our lab has been on simple particle sizing, while

Piazza and coworkers have been concerned with a number o f optical and physical

parameters.

16

Repulsion VR

VrTotal Potential Energy V-

Separationm ax

PotentialEnergy

Secondary Minimum

M agnetic Attraction

A ttraction V

Primary Minimum

Figure 1.4 The interaction param eter o f the particles which have charges, and magnetic interaction. This plot is adapted from T. Sato and R. Ruch, "Stabilization o f Colloidal Dispersions by Polymer Adsorption," p42, Marcel Dekker, Inc., NY, 1980 and R. E. Rosensweig, "Ferrohydrodynamics," p49, Cambridge Univ. Press, NY, 1985.

17

1.5 REFERENCES

1.1 T.-H., Lin and G. D. J. Phillies, J. Phys. Chem. 8 6 , 4073 (1982).

1.2 G. D. J. Phillies, Biopolymers 24, 379 (1989).

1.3 A. Vailati, D. Asnaghi, M. Giglio, and R. Piazza, Phys. Rev. E 48(4) (1993).

1.4 (a) F. Lanni, D. L. Taylor, and B. R. Ware, Biophysics J. 35, 351 (1981).

(b) T. Chang and H. Yu, Macromolecules 17, 115 (1984).

1.5 P. S. Russo, M. Mustafa, T. Cao, and L. K. Stephens, J. Coll. Inter. Sci. 122, 120 (1988).

1.6 W. Brown and R. Rymden, Macromolecules 21, 840 (1988).

1.7 G. S. Ullmann, K. Ullmann, R. M. Lindner, and G. D. J. Phillies, J. Phys. Chem. 89, 692 (1985).

1.8 K. Ullmann, G. S. Ullmann, and G. D. J. Phillies, J. Coll. Inter. Sci. 105, 315, (1985).

1.9 Z. Bu and P. S. Russo, Macromolecules 27, 1187 (1994).

1.10 (a) D. W. Pohl, S. E. Schwarz, and Imiger, Phy. Rev. Lett. 31, 32 (1973).

(b) L. Leger, H. Hervet, and F. Rondelez, Macromolecules 14, 1732 (1981).

(c) E. J. Amis, P. A. Janmey, J. D. Ferry, and H. Yu, Macromolecules 16, 441 (1983).

1.11 (a) F. Lanni and B. R. Ware, Rev. Sci. Instrum. 53(6), 905 (1982).

(b) M. R. W attenbarger and V. A. Bloomfield, Z. Bu, and P. S. Russo, Macromolecules 25, 5263 (1992).

(c) J. Davoust, P. F. Devaux, and L. Leger, The EMBOJ. 1(10), 1233 (1982).

(d) B. R. Ware, American Lab. 16, April (1984).

(e) B. A. Smith, Macromolecules 15, 463 (1982).

(f) K. Zero and B. R. Ware, J. Chem. Phys. 6 6 , 1610 (1984).

1.12 (a) T. L. James and G. G. M acDonard, J. Magn. Reson. 11, 58 (1973).

(b) W. Brown, P. Stilbs, and R. M. Johnsen, J. Poly. Sci., Poly. Phys. Ed. 20,1771 (1982).

18

1.13 D. N. Petsev and N. D. Denkov, J. Coll. Inter. Sci. 149(2), 329 (1992).

1.14 G. K. Batchelor, J. Fluid. Mech. 74, 1 (1976).

1.15 G. D. J. Phillies, G. S. Ullmann, K. Ullmann, and T. -H. Lin, J. Chem. Phys. 82, 5242 (1985).

1.16 (a) A. G. Ogston, B. N. Preston, and J. D. Wells, Proc. Royal Soc. London A.133, 297 (1973).

(b) A. R. Altenberger and M. Tirrell, J. Chem. Phys. 80, 2208 (1984).

1.17 L. Leger, H. Hervet, and F. Rondelez, Macromolecules 14, 1732 (1981).

1.18 (a) W. Brown, P. Stilbs, and R. M. Johnsen, J. Polym. Sci., Polym. Phys. Ed. 20,1771 (1982).

(b) E. O. Stejskal and J. E. Tanner, J. Chem. Phys. 42, 288 (1965).

1.19 (a) J. W. Klein and B. R. Ware, J. Chem. Phys. 80(3), 1334 (1984).

(b) B. R. Ware and W. H. Flygare, Chem. Phys. Lett. 12(1), 81 (1971).

1.20 R. J. Hunter, "Foundations o f Colloid Science," Vol. 1, Clarendon Press, 1987.

1.21 G. S. Manning, Biopolymers 11, 937 (1972).

1.22 (a) T. Odijk, Macromolecules 12, 6 8 8 (1979).

(b) J. Skolnick and M. Fixman, Macromolecules 10(5), 944 (1977).

1.23 D. B. Roitman, R. A. Wessling, and J. McAlister, Macromolecules 26, 5174 (1993).

1.24 (a) General Booklet o f Rhone-Poulenc "ESTAPOR microspheres," 1987.

(b) L. B. Bangs, "Uniform Latex Particles," Seradyn Inc., Indianapolis, IN 1984.

1.25 J. C. Diniel, J. L. Schuppiser, M. Tricot, U.S. Pat. 4358388, Eur. Pat Application 38730.

1.26 Seradyn Particle Technology newsletter Vol. 3, Spring 1992.

1.27 (a) McGraw-Hill Encyclopedia o f Science and Technology, Vol. 8 , 1977.

(b) W ebster's Third N ew International Dictionary, 1976.

1.28 A. H. Morrish, "The Physical Principles o f Magnetism," John Wiley & Sons, Inc, p361, 1965.

19

1.29 D. S. Parasnis, "Magnetism," Science Today Series, Harper & Brothers, New York, 1961.

1.30 M. Fermigier and A. P. Gast, J. Coll. Inter. Sci. 154(2), 522 (1992).

1.31 P. S. Russo, M. J. Saunders, and L. M. DeLong, Analytica ChimicaActa 189,69, (1986).

1.32 R. Piazza and V. Degiorgio, PhysicaA 182, 576 (1992).

1.33 D. Sohn, L. M. DeLong, and P. S. Russo, Mat. Res. Soc. Symp. Proc. 248, 247 (1992).

CHAPTER 2

PROBE DIFFUSION OF MAGNETIC LATEX PARTICLES

* Portions o f this chapter are taken with permission (Appendix C) from the Material Research Society Symposium Proceeding, Vol. 248, pp. 247-252 (1992). Copyright © 1992 by Materials Research Society.

20

21

2.1 BACKGROUND

Dynamic light scattering has been used for a long time to study the mobility o f

colloidal particles in complex polymeric solutions [2.1]. The most common probes are

uniform latex particles, which generally scatter much more than the polymer matrix.

W hen the probe particles are present at low concentrations and when they scatter much

more than their surroundings, the measured diffusion coefficient is that o f the probe

through the polymer matrix. A number o f issues regarding the translational motion o f

such probes, ranging from the connectivity and stiffness o f the polymer matrix to the

size and shape o f the particles, are still being pursued in this and other laboratories.

Departing somewhat from these concerns, the present report is a preliminary account

o f rotational motion o f spherical particles in a polymer matrix, as determined by

dynamic light scattering. This is only possible if the spherical particles possess optical

anisotropy. Commercially available magnetic latex particles [2 .2 ] are a suitable choice,

due to the crystalline nature o f their magnetite (Fe3 0 4) inclusions.

The light scattering measurements in this report were conducted in Vv and H v

geometries. The first letter refers to the polarization sense o f the detected light. The

second, smaller letter refers to the polarization sense o f the incident light. Horizontal,

H, refers to the light being polarized in the scattering plane defined by the incident and

detected rays. Vertical light is polarized perpendicular to the scattering plane (Figure

2 . 1).

The intensity-intensity autocorrelation function o f scattered light in the familiar

[2.3] homodyne experiment has the form, G ^ ( t ) = B(1 + f(A) | g ^ (t) |2), where g 1 (t)

is the electric field autocorrelation function. For optically isotropic, geometrically

spherical particles such as most latex spheres, g ^ 1 Vv ( 0 is an exponential with decay

rate

r V v = q 2 Dt (2.1)

22

J *n j r

Sam ple A nalyserfTemp. Control)

Amplifier / Dlscrlmlnalor

p m r

CorrelatorLangjey Ford Model 1096

Figure 2.1 Light scattering alignment for the H v and Vv measurement.

23

where q = 4 7 t. n .sin(0/2)/A,o (n = solution refractive index, 0 = scattering angle, X0=

laser wavelength in vacuo) and Dt is the translational diffusion coefficient.

Optically anisotropic particles are, in general, not simple [2.4], However, the

case o f cylindrical symmetry (both optical and geometric) is well understood, especially

in the absence o f coupling between translational and rotational motion [2.3], Spinning

about the thin axis o f such rod-shaped particles is generally not visible but end-over-

end tumbling, characterized by the rotational diffusion coefficient Dr, is. The

appropriate correlation function is g^HyOO QC P2 exp(-q 2 Dt - 6 D r)x, where 3 is the

difference between polarizability in parallel and perpendicular directions with respect to

the cylinder axis. The H v decay rate is therefore

This expression is also valid for a hypothetical spherical particle with a

cylindrical, optically anisotropic inclusion in its center, as shown in Figure 2.2. This is

a starting model for our magnetic latex particles, although its suitability may only be

judged by comparison to experiment. In real particles, there is no reason to assume

that just one magnetic inclusion is located at the geometric center o f the particle;

indeed, according to the patent literature and our own electron microscopic

observations, this seems unlikely. We appeal to the model depicted in Figure 2.2

primarily for its simplicity.

Translational diffusion o f a sphere o f radius R in a simple solvent o f viscosity r\

0 is governed by the Stokes-Einstein relation, Dt = kT/(6 7 tri0 R), where kT is the

thermal energy per molecule. In complex solutions, this equation has been rearranged

to obtain a "microviscosity" associated with translational motion,

r Hv = q2°t + 6 D r (2 .2)

1 —

^ ~ 67tD,R(2 .3)

24

Figure 2.2 Idealized model o f a magnetic latex particle (not to be confused with actual latex particles).

25

where R is usually measured from dilute binary solution in a simple solvent. In general,

there is no reason for to equal the macroscopic solution viscosity r\ because a small

particle can "slip between" polymer molecules that comprise the matrix without the

wholesale reorganizations that attend imposition o f a macroscopic shear gradient.

Intuitively, one might expect the microviscosity to lie between the solution and solvent

viscosities. In fact, large differences between t | and r)1̂ have been reported [2.5, 2.6].

Rotational diffusion o f a sphere in a simple solvent is given by

kTr 87CTIJ13

° r = ^ ^ 3 (2-4)

This can be adapted to yield a rotational microviscosity, representing the opposition

provided by the solution to rotation o f the probe

r kT ^Tin = „ ^ (2 5)

^ 87iDrR

We are not aware o f any other experimental measurements o f the retardation o f

rotational diffusion o f a submicron spheres by a polymer matrix.

2.2 EXPERIMENT

Latex particles o f advertised diameter 700 nm were purchased from Seradyn

Particle Technology Division. According to the manufacturer, they consist o f 58%

Fe30 4/42% polystyrene by weight; we confirmed this by thermogravimetric analysis.

Approximately 5 x 10 ' 5 g/ml Seradyn latex/water solution was prepared by adding dust-

free, 18M n-cm resistivity w ater from a Bamstead Nanopure purifier equipped with a

spiral wound ultrafilter to a stock solution prepared with dust free water, centrifuged to

remove very large particles, and passed through a well rinsed 600 nm diameter

polyester filter (Nuclepore). This effectively isolates the smaller particles in the sample

distribution.

26

A smaller magnetic latex, advertised diameter 54 nm, was obtained from

Polysciences Inc. Thermogravimetric analysis showed these particles to be 12%

Fe 3 0 4 /88% polystyrene by weight. Latex stock solution, 6.25X10"4 g/ml, was prepared

similarly to the Seradyn sample except that a 0.22 pm Durapore filter (Millipore) was

used. This does not significantly fractionate the particles, but only reduces the dust

level.

Polystyrene sulfonate, sodium salt (NaPSS: advertised molecular weight

500,000) was purchased from Scientific Polymer Products Inc. Stock solution was

prepared at 0.05 g/ml with dust-free water, and centrifuged 12 hours at 4700 g.

Dilutions were prepared by addition o f dust-free water and magnetic latex stock.

The light scattering device and the general procedures are detailed elsewhere

[2.4], The Seradyn sample was measured at 30 ± 0.1°C, X0 = 632.8 nm. Polysciences

magnetic latex and ternary systems (Magnetic latex/NaPSS/W ater) were obtained from

four different angles at 25 ± 0 .1°C and X0 = 488.0 nm.

The macroscopic viscosities o f ternary systems were measured on a Brookfield

LVTDCP cone and plate viscometer (range, 0.5 - 1,000 centipoise, cP). A t the low

concentrations and molecular weight o f NaPSS explored so far, shear thinning was

very minor in the available shear rate range (75 - 450 Hz); straightforward

extrapolation to zero shear rate was made.

2.3 RESULTS AND DISCUSSION

2.3.1 IN DILUTE BINARY SOLUTION

Figure 2.3 shows the q-dependence o f decay rate for Seradyn magnetic latex in

pure water. As expected, the slopes o f the H v and Vv plots are similar, and the H v plot

has a finite intercept. One may obtain three estimates o f the particle radius from this

plot. The most conventional one, R Sw (82.8 ± 0.5 nm), comes from the diffusion

coefficient obtained as the slope (hence, the superscripts) o f the V v plot, Eq. 2.1, plus

27

the Stokes-Einstein relation. RsHv (98 ± 4 nm) is similarly obtained, except the

difliision coefficient is the slope o f the H v plot; see Eq. 2.2. The final estimate, RTHv

(110 ±20 nm), comes from Eq. 2.4 where Dr is one-sixth the intercept (hence,

superscript I) o f the H v plot; see Eq. 2.2. The reasonable agreement among these three

values reflects the overall adequacy o f the simple model depicted in Figure 2.2 for this

sample. However, electron microscopy shows not one large magnetic inclusion located

in the center o f the particle, but many small inclusions distributed more or less

symmetrically about the center. This appears to be almost indistinguishable from the

model in Figure 2.2 by dynamic light scattering, and it seems that these particles rotate

with minimal friction about a polar axis, without the wobbling that would result if one

or more heavy magnetite inclusions were asymmetrically located, causing imbalance. It

is emphasized that particles "just out o f the bottle" do not yield such exemplary

behavior, filtration and centrifugation o f the suspension are first required to isolate the

small particles at the tail end o f what electron microscopy shows to be a wide size

distribution.

The Seradyn latex particles have most o f the characteristics desirable in

rotational probe diffusion experiments. However, the H v intercept is rather close to

zero, given realistic uncertainties o f 10 - 20%. The accuracy with which D r could be

determined would become poor as matrix polymer was added. Attem pts to make H v

measurements directly at zero scattering angle were complicated by a very slowly

decaying component, possibly due to weak aggregation o f the particles in the

geomagnetic field (as described below, large aggregates are very evident in an applied

magnetic field; presumably smaller aggregates induced by the Earth's field might

contribute strongly to the scattering at 0 = 0). So rather than pursue the Seradyn

spheres as probes, the smaller particles from Polysciences were tested for suitability in

conventional finite angle Vv and Hv experiments similar to those shown in Figure 2.3.

28

4 0 0 0

1 1 0 nm (in tercep t)

99 nm (slope)3000-

2000 - 83 nm

r e

1000

2 / 1010cm-2

Figure 2.3 Decay rates from H v and Vv measurements vs. q2, for Seradyn magnetic latex in pure water. Apparent particle radii are indicated, along with linear correlation coefficients for the least squares lines.

29

For the Polysciences spheres, agreement among the three available radii was

poor (RSvv = 23.6 ± 0.2 nm, in close agreement with the manufacturer's claim; RsHv =

134 ± 29 nm; R!Hv = 43 ± 1 nm). N ote that the H v radii are 2 and 6 times larger than

the Vv value (Figure 2.4).

Electron micrographs reveal some minor size heterogeneity. Perhaps the larger

particles exhibit greater depolarized scattering and are weighted more heavily in the H v

measurements. But the difference between the slope- and intercept-determined radii

from the same H v data set suggests other factors. The low magnetite content for the

Polysciences spheres, coupled with their small size, results in relatively few inclusions

per particle, favoring non-symmetric distribution. M agnetite particles appear as dark

spots in transmission electron micrographs, and do appear to be asymmetrically located

in the Polysciences spheres (typically, near the periphery as also noted in Ref. [2.2]).

The density o f magnetite, 5.2 g/cm [2.7], greatly exceeds that o f polystyrene (~ 1.05

g/cm 3 in latex form). So it is not unreasonable that rotation o f these particles resembles

more a "wobble" than a "spin". The friction associated with such motion should be

higher than for rotation o f a sphere about its midpoint, leading to high radius values

when Dr is interpreted via Eq. 2.4. M oreover, Eq. 2.2 from which Dr is measured is no

longer strictly correct if the optical and friction axes do not coincide. Nevertheless, we

shall continue to use it to obtain apparent Dr values which may approximately represent

"rotation" o f these unusual particles.

2.3.2 IN P R E SE N C E O F PO LY (STY R EN ESU LFO N A TE)

To date, only the Polysciences particles have been used in probe diffusion

measurements; despite possible imbalance and the resulting ambiguities, their small

average size results in a large H v decay rate in simple binary solution, making it easy to

observe rotational slowing as NaPSS is added.

30

20000

40 nm (intercept)

100 nm (slope)

1 5 0 0 0 -

I

10000 -

5 0 0 0 -24nm

10 126 80 2 4

q2/1010 cm"2

Figure 2.4 Decay rates from H v and Vv measurements vs. q2, for Polysciences magnetic latex in pure water. Apparent particle radii are indicated, along with linear correlation coefficients for the least squares lines.

31

However, the depolarized signal from the small particles is much weaker than the

Seradyn spheres, reflecting lower magnetic content.

Longer acquisition times can partially compensate for this; at each concentration, about

one day was required to obtain H v measurements at several angles. D t was obtained as

the slope o f rVv vs. q2 plots, and D r from the intercept o f rHv vs. q2 plots (Figure 2.5

and 2.6). Figure 2.7 shows the microviscosities, together with the macroscopic

solution viscosity; was computed using RsVv, while r)1̂ was computed using R rHv.

All viscosities increase with concentration, but the microviscosities are evidently

smaller than the macroscopic viscosity. This precludes strong specific interaction o f

the polymer and latex. Indeed, none is expected, since both contain mutually repulsive

negatively charged sulfate groups. It appears that treating the polymer solution as a

Stokes-Einstein continuum is invalid for either rotational or translational motion in this

case. Direct comparison between r^ a n d is not straightforward because o f the

possible imbalance o f the particles.

2 .4 S U M M A R Y

Dynamic light scattering is sensitive, but not extremely so, to structural

differences in magnetic latexes. Thus, the imbalance o f the Polysciences spheres was

clearly evident (well before confirmation by electron microscopy) while the Seradyn

magnetic spheres could be fitted to the model shown in Figure 2.2, even though

electron micrographs reveal many small, symmetrically located inclusions instead o f one

central magnet. The Seradyn latex would make a suitable probe o f rotational motion if

the intercept o f THv vs. q2 plots could be determined somewhat more reliably. It is

possible that improved uniformity, cancellation o f the geomagnetic field, or use o f latex

spheres with nonmagnetic, optically anisotropic inclusions, will prove useful. The

unbalanced nature o f the smaller Poly sciences magnetic latex spheres causes some

ambiguities in the interpretation.

32

No NaPSS 2.4xl0 - 3 g/ml NaPSS 4.5xl0 ' 3 g/ml NaPSS 8.3x1 O' 3 g/ml NaPSS

10000

8000

6000

4000

2000

0

0 4 6 8 10 122

q2 / 1010 cm-2

Figure 2.5 Decay rates from Vv measurement vs. q2 for Polysciences magnetic latex particles in vaiying NaPSS concentration.

33

□ No NaPSS

O 2.4x10'3 g/ml NaPSSA 4.5x10 ' 3 g/ml NaPSS

V 8.3xl0‘3 g/ml NaPSS

20000 -

1 5 0 0 0 -

I

10000 -

5 0 0 0 -

0 2 4 6 8 10

q2 / 1010 cm"2

Figure 2.6 Decay rates from H v measurement vs. q2 for Polysciences magnetic latex particles in varying NaPSS concentration.

34

5

4

r)/ cP3

2

1

0

0 .

Figure 2.7 M acroscopic shear viscosity o f NaPSS solution, plus rotational and translational micro-viscosity for Polysciences magnetic latex.

□ V iscosity (zero shear rate) D

O T ransla tional M icroviscosity

A R otational M icroviscosity

a

A

o

A

O

A

O

1 1 1 1 1 1 1------------------ 1—

000 0 .002 0 .004 0.006 0.008

C N a P S S ^ 1

35

Analysis via expressions appropriate for the single encapsulated magnet model

produced rotational microviscosities similar to the translational values, in the range o f

NaPSS concentration explored so far.

I f a dilute solution containing magnetic latex is placed above a simple magnetic

stirring apparatus (but not in physical contact with it, to avoid vibrations) and

illuminated by a laser beam, one observes that the scattered light fluctuates according

to the position o f the rotating magnet. Autocorrelation experiments, such as the one

shown in Figure 2.8, confirm that the clearly visible intensity fluctuations at low

magnetic oscillation frequencies persist to much higher frequencies, even though they

are no longer visible to the naked eye. These fluctuations are due to reorientation, or

partial reorientation, o f chain like aggregates (easily observed in a microscope) o f the

particles induced by the magnetic field. This suggests that light scattering could be

used to measure the frequency response in reorientation for such chain-like structures.

36

O

<N

" o

1.8

1.7

1.6

0 50 100 150 200 250 300

Channel Number 0.0007 s/ch

Figure 2.8 Intensity autocorrelation function (Vv) o f Seradyn magnetic latex in the presence o f an externally applied, oscillating magnetic field. The initial diffusion-like decay is followed by sinusoidal oscillations.

37

2.5 REFERENCES

2.1 (a) G. D. J. Phillies, J. Phys. Chem. 93, 5029 (1989).

(b) W. Brown and R. Rymden, Macromolecules 19(12), 2942 (1986).

(c) J. Newman, N. M roczka, and K. L. Schick, Biopolymers_28 , 655 (1989).

(d) P. S. Russo, M. Mustafa, T. Cao, and L. K. Stephens, J. Coll. Inter. Sci. 122(1), 120 (1988).

2.2 J. -C. Daniel, J. -L. Schuppiser and M. Tricot, U.S. Patent No. 4,358,388(4 Nov. 1982).

2.3 B. Berne and R. Pecora, "Dynamic Light Scattering," Wiley, N ew York, 1976.

2.4 P. S. Russo, M. J. Saunders, L. M. DeLong, S. K. Kuehl, K. H. Langley and R.W. Detenbeck, Anal. Chem. Acta 189, 69 (1986).

2.5 T.-L. Lin and G. D. J. Phillies, Macromolecules 17, 1686 (1984).

2.6 G. S. Ullmann, K. Ullmann, R. M. Lindner and G. D. J. Phillies, J. Phys.Chem 89,692 (1985).

2.7 M erck Index, Ed. 10, No. 3974, ferric oxide, M erck & Co. Inc., Rahway, NJ, 1983.

CHAPTER 3

INTERACTION BETWEEN POLYELECTROLYTES AND MAGNETIC LATEX PARTICLES*

* Portions o f this chapter are taken from a draft manuscript "Light Scattering Study o f Magnetic Latex Particles and Their Interaction with Polyelectrolytes" submitted 10/04/94 to Journal o f Colloids and Interface Science. The authors, in order, are: D. Sohn, P. S. Russo, A. Davila, D. S. Poche', and M. L. McLaughlin.

38

39

3.1 INTRODUCTION

Magnetic Latex Particles (or M LP) are commonly used in various bioseparation

techniques [3.1], Small, superparamagnetic crystallites o f magnetite give the particles

unique magnetic and optical properties. The magnetic interaction between the particles

is negligible in aqueous solution if there is no externally applied magnetic field. Even

though the optical properties, particularly the depolarization o f scattered light, differ

from those o f ordinary latex particles, the interparticle interaction is similar to that o f

common non-magnetic latex particles until a magnetic field is applied to align the spins

in the superparamagnetic inclusions. Then a strong dipolar interaction develops and the

M LP aggregate as chains that are readily pulled towards the magnet. This "chaining"

interaction, and indeed all effects o f an applied magnetic field [3.2], are beyond the

scope o f the present study.

Initial interest in the study o f optically anisotropic colloids such as the present

MLP was peaked by "probe diffusion" experiments [3.3] where the diffusion o f

colloidal particles is monitored in an attempt to learn the local structure o f some

complex polymeric matrix, such as an entangled solution or gel. These studies are

commonly performed by dynamic light scattering, however one o f the complexities o f

these systems is that the matrix may itself scatter. W ith an optically anisotropic probe,

matrix scattering would be almost completely excluded in depolarized light scattering

experiments, even in aqueous solutions where it is not possible to match [3.4] the index

o f refraction o f the solvent and matrix component. Although the M LP are fairly large

probes, their rotational m otion could provide information o f a new kind on a very local

scale. For example, non-continuum dynamical behavior might be observed in the

rotational motion, despite the relatively large size o f the M LP probes.

For the matrix, a polyelectrolyte with a negative charge was selected to prevent

the kind o f surface adsorption seen with nonionic polymers [3.5], It will be shown that

40

adsorption o f the polymer to the M LP is not significant. The extra complication o f a

polyelectrolyte matrix is, however, not trivial. Thus, while a preliminary investigation

[3.6] o f solutions without added salt quickly provided some o f the probe diffusion

information originally sought, including significant failure o f continuum hydrodynamic

relationships for translational and rotational diffusion, there are equally interesting

questions o f colloid stability and specific colloid-polymer interaction. The present

chapter is devoted primarily to these.

3.1.1 COLLOID AND POLYMER INTERACTIONS

The behavior o f charged particles has been studied by Pusey [3.7], Schaefer et

al. [3.8] by dynamic light scattering. The interaction between spherical charged

particles is commonly interpreted via the DLVO (Deijaguin and Landau, Verwey and

Overbeek) theory [3.9] which combines the electrostatic repulsion and van der Waals

attraction between the double layers. These electrical double layers are veiy sensitive

to the ionic strength o f the system. Increasing the ionic strength significantly

diminishes the thickness o f the double layer [3.10, 3.11], and the van der Waals

attraction becomes dominant. So, an electrostatically stabilized dispersion is

aggregated when the ionic strength o f the dispersion medium is raised sufficiently.

There continues to be serious disagreements as to the nature o f the interaction among

charged particles at very low ionic strengths [3.12], centered on the issue o f whether or

not there may exist an attractive interaction to counterbalance the electrostatic

repulsion.

The added polymer chains present in this study could, in principle, stabilize or

destabilize the colloidal particles [3.13, 3.14, 3.15], Colloid-polymer interactions can

be divided into "bound polymer" and "free polymer" types. Polymers bound to the

latex particle can impart stability through a steric mechanism (preventing the close

approach o f the latex particles) or induce aggregation by "bridging" tw o or more latex

41

particles. The best understood interaction between free polymers and colloidal

particles is depletion flocculation, first identified by Asakura and Oosawa [3.16], This

effect arises due to the osmotic force acting on two colloidal particles as a result o f the

exclusion o f polymeric material from the zone between them when they are separated

by a small distance. That is, when two particles are separated by a distance comparable

to the radius o f the added polymer, there is insufficient space for the polymer, so it

evacuates the region leaving only solvent in the depletion layer. Then the osmotic force

o f the polymers on the outer surfaces o f the colloidal pair drives the particles closer

together. Feigin and N apper [3.17] proposed a depletion stabilization mechanism by

considering two particles at a somewhat greater distance. As long as the colloidal

particles are separated by a distance greater than the polymer diameter, such that the

polymer can still exist between them without serious loss o f configurational entropy,

the bulk solution will resist exclusion o f polymer from that region, thereby providing a

degree o f stabilization. This resistance is based on polymer-polymer repulsions

(excluded volume or electrostatic) and increases with bulk concentration for polymers

in a good solvent. Polyelectrolytes [3.18], especially NaPSS, in salt solution have been

studied by Koene & Mandel [3.19] and Drifford & Dalbiez [3.20], It would be

expected that, in general, polyelectrolytes would impart stability by a combination o f

electrostatic and steric mechanisms [3.21], Buscall [3.22] developed a qualitative

approach to the steric stabilization with polyelectrolytes in a colloid system.

3.1.2 STATIC AND DYNAMIC LIGHT SCATTERING

Static light scattering can reveal the presence o f clusters or aggregates. The

angular dependent intensity obeys I(q)/I(0) « 1 - q2 Rg2/3 + ■■• where I is the scattering

intensity at a given scattering vector, whose magnitude is q = 4 7 r.n.sin(0 / 2 )A o (n =

refractive index; 0 = scattering angle; \ Q - incident light wavelength in vacuo). It is

easily seen that the radius o f gyration, Rg, is obtained in the limit o f low q as:

42

Rg = (3<flnl(q)/rfq2 ) 1 /2 (3.1)

This Guinier relationship [3.23] is valid at suitably low concentrations.

Dynamic light scattering measurements were conducted in V v and H v

geometries. The first letter refers to the polarization sense o f the detected light. The

second, smaller letter refers to the polarization sense o f the incident light. Horizontal,

H, refers to the light being polarized in the scattering plane defined by the incident and

detected rays. Vertical light is polarized perpendicular to the scattering plane.

The intensity autocorrelation function o f scattered light in the familiar [3.24]

homodyne experiment has the form, ( t) = B(1 + f(A) | g^'l (x) |2), where g ^ ( T) is

the electric field autocorrelation function. For monodisperse, optically isotropic,

geometrically spherical particles such as most latex spheres, g ^ 1 Vv (x) ' s an exponential

with decay rate

where Dt is the translational diffusion coefficient. For polydisperse scatterers or

scatterers that exhibit internal relaxation modes, including rotation o f nonspherical

particles, r Vv often increases a little faster than q2. In these cases, D t is more

The dynamic light scattering from optically anisotropic particles is, in general,

not simple [3.25], However, the case o f cylindrical symmetry (both optical and

geometric) is well understood, especially in the absence o f coupling between

translational and rotational motion. Spinning about the thin axis o f such rod-shaped

particles is generally not visible but end-over-end tumbling, characterized by the

rotational diffusion coefficient Dr, is. The appropriate correlation function is g*-1 Vv(T)

oc p 2 [exp(-q2 D t - 6 Dr)x], where p is the difference between polarizability in parallel and

r Vv = q2 Dt (3.2)

accurately obtained as the zero-q limit o f Dapp = r Vv/q2-

(3.3)

43

perpendicular directions with respect to the cylinder axis. The H v decay rate is

therefore:

THv = q 2 Dt + 6 Dr (3.4)

Equation 3.4 will be appropriate for anisotropic inclusions in a spherical particle if they

are uniformly distributed and isotropically arranged.

Translational diffusion o f a sphere o f radius R in a simple solvent is governed by

the Stokes-Einstein equation,

D t = kT / 6 7 ir |0R (3.5)

where r | 0 is the solvent viscosity, k is Boltzmann's constant, and T is the absolute

temperature. In complex solutions, Equation 3.5 has been rearranged [3.26] to obtain

a "microviscosity" associated with translational motion,

= kT / 67iDtR = ft/(67tR) (3.6)

where R is usually measured from dilute binary solution in a simple solvent and ft is the

friction factor associated with translational motion (we assume zero probe

concentration, so there is no difference between self and mutual friction factors). In

general, there is no reason for q 1̂ to equal the macroscopic solution shear viscosity q

because a small particle can "slip between" polymer molecules that comprise the matrix

without the wholesale reorganizations that attend imposition o f a macroscopic shear

gradient. Intuitively, one might expect the microviscosity to lie between the solution

and solvent viscosities. In fact, large differences between q and q 1̂ have been reported

[3.27] though there is some disagreements exist as to the magnitude, even in similar

systems.

The rotational diffusion o f a sphere in a simple solvent is given by

Dr = kT / 87tq0 R 3 (3.7)

W e can obtain from the foregoing equations various radii o f the particles.

R sVv = kT /e itrioD t (3.8)

44

R*Hv = ( k T /S T n ^ ) 1'3 (3.9)

R sHv = k T / e ^ i ^ (3.10)

where RVv is the radius from Vv measurements extrapolated to zero q, and R1̂ and

R sHv are radii from the intercept and initial slope o f H v measurements, respectively. It

is understood that D t in Equation 3.10 comes from the H v measurement, and may have

a different value than D t in Equation 3.8.

3.2 EXPERIMENT

W ater was obtained from a Bam stead Nanopure 5-stage purifier including a

spiral wound ultrafilter (50 A), which is in turn fed by a Millipore Milli-Q three-stage

system. The measured resistivity at the tap is 18 M Q-cm. The w ater is absolutely free

o f dust or other impurities, as determined by visual inspection o f the light scattered

from the water when illuminated by a tightly focused argon ion laser beam at lOOx

magnification using a custom made light scattering spectrometer. The light scattering

cells were treated with chlorotrimethylsilane, resulting in a hydrophobic, methylated

surface on the inside to facilitate rising, and tested for cleanliness by visual inspection

as just described for w ater testing.

Superparamagnetic latex particles have variable number o f magnetite (Fe3 0 4)

inclusions approximately 1-20 nm in size. Three samples were purchased from

different sources; they are referred to by the vendor name. Steps were taken to clean

the particles and reduce size polydispersity. Seradyn latex particles o f advertised

diameter 0.7 pm were purchased from Seradyn Particle Technology Division.

According to the manufacturer, they contain 58% Fe3 0 4 / 4 2 % polystyrene by weight;

this was confirmed by thermogravimetric analysis. Approximately 5 x l0 ’ 5 g/ml Seradyn

latex/water solution was prepared by adding dust-free water. The sample was

centrifuged to remove very large particles, and passed through a well rinsed 0 . 6 pm

45

diameter polyester filter (Nuclepore). This effectively isolates the smaller particles in

the sample distribution.

A smaller magnetic latex, advertised diameter 0.054 pm, was obtained from

Polysciences Inc. By thermogravimetric analysis, these particles contained 12%

Fe3 0 4 / 8 8 % polystyrene by weight. Latex stock solution, 6.25X10-4 g/ml, was prepared

similarly to the Seradyn sample except that a 0.22 pm Durapore filter (Millipore) was

used. This does not significantly fractionate the particles, but only reduces the dust

level.

The third magnetic latex particle containing 6 8 % iron oxide by weight was

purchased as a 2% suspension from Bangs Laboratories, Inc. The surface sites o f these

particles are carboxyl groups (advertised surface charge 197 peq/g and parking area

2.7 A2/COOH group). The advertised radius was 0.4 pm. A 0.055 wt% solution was

prepared and filtered with a well-rinsed Millipore 0.45 pm Millipore H A membrane to

make a stock solution containing particles with average radius 0.10 pm. This is a

convenient size; smaller magnetic latex particles often have such a w eak depolarized

signal that very long acquisition times are needed to collect data o f high quality, while

larger latex particles rotate too slowly for convenient measurement on the linearly

spaced, 4-bit Langley-Ford 1096 autocorrelator that was used.

Polystyrene sulfonate sodium salt (NaPSS; advertised M = 70,000) was

purchased from Scientific Polymer Products Inc. This NaPSS is not the narrow

distribution standard material, but GPC light scattering in 0.01 M NaCl shows the

polydispersity is not excessive: M^/Mj, < 1 .6 and Mw = 70,000 as claimed by the

supplier. O f special importance to the present study is that elemental analysis (by

Oneida Reseach Service, Inc.) shows that the sample is «100% sulfonated. A stock

solution at 0.0428 g/ml NaPSS/HzO was prepared and filtered with a Millipore 0.1 pm

M illex -W filter. Salt solutions at 0.01 M, 0.05 M, and 0.1 M NaCl solution were

46

prepared with GR grade (EM Science) NaCl and filtered with a 0.1 pm M illex -W

Millipore filter.

To make the ternary M LP/NaPSS/H20 solution, 1 ml o f H20 was put into each

o f 4 silanated cells, and 0.01 ml, 0.05 ml, 0.1 ml, and 0 . 2 ml NaPSS stock solution

were added to the 4 cells, respectively. Then, 0.2 ml o f magnetic latex stock solution

was added to each o f the cells. The final latex concentration was 1,4x lO^/o; apparent

translational diffusion coefficients ranged from 2.7-2 .82xl0 '8 cm " 2 sec" 1 between

lx l0 '3% and lx l0 '5% range o f concentrations. Figure 3.1 shows the magnetic latex

particle concentration dependent T vs. q2 plot. These concentrations are a little lower

than ideal. Low concentrations can enhance the significance o f number fluctuations

[3.28] relative to the desired diffusive decay o f the correlation function. Optical

settings in the detector had to be adjusted to measure a somewhat larger volume than

normally used to keep the number fluctuation problem at bay. The above procedure

results in a 2 0 % spread o f the final latex concentration, which is insignificant at the low

total latex content used. The final NaPSS concentration o f each cell was 0, 7.016x1c4,

3 .292xl0 '3, and 6 .114xl0‘3 g/ml. The quaternary solutions M LP/NaPSS/H 2 0/N aC l

solutions at 0.01 M, 0.05 M, and 0.1 M NaCl were prepared by the same process

except for changing the water to sodium salt solution. The quoted salt concentrations

refer to added salt. At zero added salt, the conductivity at the highest NaPSS

concentration was the same as that o f a 0.004 M NaCl solution. Dialysis reduced this

by half, so we estimate the residual salt concentration as < 0.002 M.

The custom-built light scattering spectrometer uses a Lexel Model 95 argon ion

laser capable o f producing about 1 W at 488.0 nm. The laser beam was approximately

vertically polarized, so a Glan-Thomson polarizer o f low-power rating (Karl Lambrecht

M GT3E5; 1W cm"2) was sufficient to eliminate the small horizontal component.

47

2.5

□ 2.2x10^% MLP

O 5.5x10"4% MLP

A 1.1x10'3%MLP

V 2.2x10'3% MLP

O I.IxKTVo MLP2 .0 -

0 .5 -

0.06 80 2 4

q2 /1010 cm’2

Figure 3.1 M agnetic latex particle concentration dependent decay rate vs. q2.Apparent translational diffusion coefficients did not change much between lx l0 '3% and l x l 0 '5% range o f concentrations.

48

A nicol rated for high power (Karl Lambrecht M GLQD 8 500 Wcm"2) served as the

analyzer and can almost fully extinguish the direct laser beam in a zero scattering angle

alignment (not used here). Each polarizer was mounted in a New port Research Model

470-B rotator, with angular resolution to 0.0012°. The Hamamatsu R928P

photomultiplier tube was connected to a Pacific Precision Instruments Model AD 126

photon counting system. The correlator was a 272-channel Langley Ford Model 1096.

Samples were supported in a refractive index matching toluene bath. Temperature was

controlled to 25 ± 0.1 °C by nonturbulent water circulation through a pure copper

block. Further details o f the scattering device, alignment and safety precautions appear

elsewhere [3.25],

Each correlation function was gathered as a series o f short runs, and each short

run was analyzed by second-order cumulants [3.29], After discarding "outliers" the

remaining functions are summed and reanalyzed by first-third order cumulants. Thefj

average decay rate, T, was plotted vs. q~ for six different q values (corresponding to 0

= 30°,45°,60°,90°, 120°, and 135°). And then the diffusion value from 17q^ is

compared with T/q^ vs. q^ plot to consider the deviation from the 17q^ plot. For the

SLS experiment the total intensity o f the 16 different samples was measured every 10°

starting from 20° to 120°, and the first 7 points were analyzed.

3.3 RESULTS AND DISCUSSION

3.3.1 INTERPARTICLE INTERACTIONS IN SALT CONDITION

3.3.1.1 PARTICLE CHARACTERIZATION IN PURE WATER

Figure 3.2 shows typical decay profiles for Bangs MLP in Vv and H v

geometries. Though not perfect single exponential, there is no difficulty extracting the

mean decay rate from such data via cumulants analysis. The sudden leveling that slow

number fluctuations would cause is not evident. Figure 3.3(a), (b), and (c) shows the

decay rates from Vv and H v DLS experiments for three different particles in pure water.

49

0

1

Vv

■2Hv

3

•40.0 0.2 0.4

t / 1 0 " 2 sec

Figure 3.2 Typical decay profiles for Bangs magnetic latex particles in w ater in Vv and H v geometries. Measurements have been done at 60° scattering angle and at 25°C.

50

Vv

S eradyn M L P

3.0

t—h

oxx.

I-H Vvi r

i2 / 1 0 1 0 c m ' 2

Figure 3.3(a) Seradyn magnetic latex particles' (1) decay rate vs. q2 plot o f Vv and H v measurement. (2 ) T/q2 vs. q2 plot.

51

P olysciences M L P

q2/1 0 10 cm'2

Hv1 5 -

Vv

108640 2

q2/1010 cm'2

Figure 3.3(b) Polysciences magnetic latex particles' (1) decay rate vs. q2 plot o f Vv and H v measurement. (2) I7q 2 vs. q2 plot.

52

cn

1=0

oo1O

2 . 8 -

Vv

Bangs MLP2.4-

&Q

2.04 6 8 100 2

q2/1010 cm'2

Hv

Vv-21

6 8 100 2 4

q2/ 1010 cm-2

Figure 3.3(c) Bangs magnetic latex particles' (1) decay rate vs. q2 plot o f Vv and H v measurement. (2 ) T/q2 vs. q2 plot.

53

The decay rate was determined by 3rd cumulant fitting except for the small MLP o f

Figure 3.3(b). In this case, a single exponential nonlinear least squares fit with floating

baseline was used to reduce the effects o f baseline uncertainty. For the Seradyn MLP,

the slopes o f the H v and Vv plots are similar, and the H v plot has a finite intercept.

Closer inspection does reveal some upward curvature in the Vv plot, probably due to

polydispersity (as the scattering angle is increased, scattering from the larger particles

decreases). The upward curvature is reflected in a positive slope for the D app vs. q2

plot (Figure 3.3(a)). Using Equations 3.8-10, one may obtain three estimates o f the

particle radius from the combined H v and Vv dynamic light scattering; these radii

appear in Table 3.1. The reasonable agreement among the three values for the Seradyn

M LP reflects the overall adequacy o f a simple model which has one magnetite inclusion

located symmetrically in the center o f the latex. However, electron micrographs

(Figure 1.2) shows not one large magnetic inclusion located in the center o f the

particle, but many small inclusions distributed more or less symmetrically about the

center and predominantly near the particle surface, in agreement with Daniel et al.

[3.30], A structure with many inclusions symmetrically located near the surface

appears to be almost indistinguishable by dynamic light scattering from the simple,

single symmetric inclusion model. It is emphasized that particles "just out o f the bottle"

do not yield such exemplary behavior; the filtration and centrifugation steps described

above are required to isolate the small particles at the tail end o f what electron

microscopy shows to be a wide size distribution.

For the Poly sciences spheres, agreement among the three available radii was

poor (Table 3.1 and Figure 3.3(b), R sVv = 23.6 ± 0.2 nm, RsHv = 134 ± 29 nm; R 1̂ = 43

± 1 nm). The H v radii are 2 and 6 times larger than the Vv value. Electron

micrographs reveal some minor size heterogeneity.

54

Table 3.1 Dynamic light scattering data o f three different magnetic latex particles in pure water. Dynamic radii are from slope o f Vv (RsVv), H v ( R sh v) . and intercept from H v (R!Hv). Unit: A.

R Vv r SH v r IH v Surface

Function

% Fe3 0 4

Seradyne

Lot # 1125 M l-070/60/319

830±5 980±40 1 1 0 0 + 2 0 0 -so3 58.0%

Polysciences

Lot #404206

234±2 1340+290 430±10 -so3 1 2 .0 %

Bangs

Lot #L910402DM l-070/60/380

1030± 10 1100+30 1110±490 -COOH 67.6%

55

Perhaps the larger particles exhibit greater depolarized scattering and are weighted

more heavily in the H v measurements, but the difference between the slope- and

intercept-determined radii from the same H v data set suggests other factors. The low

magnetite content for the Polysciences spheres, coupled with their small size, results in

relatively few inclusions per particle, favoring non-symmetric distribution. M agnetite

particles appear as dark spots in transmission electron micrographs (Figure 1.2) and do

appear to be asymmetrically located in the Polysciences spheres. The density o f

magnetite, 5.2 g/cm, greatly exceeds that o f polystyrene (~ 1.05 g/cm 3 in latex form).

So we originally thought that rotation o f these particles resembles more o f a "wobble"

than a "spin." We reasoned that the friction associated with such motion should be

higher than for rotation o f a sphere about its midpoint, leading to high radius values

when Dr is interpreted via Equation 3.7. This was the assessment that appeared in our

preliminary report [3.6], It has since been pointed out [3.31] that such "wobbling"

effects are ballistic in nature. Strongly damped rotational Brownian motion should not

reflect such short-time effects. M oreover, as we pointed out originally, Eq. 3.7 from

which Dr is measured is no longer strictly correct if the optical and friction axes do not

coincide. The discrepancies in the Vv and tw o H v results are not well understood.

The refined Bangs particles were comparable in size to those in the Seradyn

sample. The slopes o f the H v and Vv plots were similar (Figure 3.3(c)) and , as for the

Seradyn sample, the three radii are in reasonably good agreement (Table 3.1).

3.3.1.2 INTERPARTICLE INTERACTIONS: NO ADDED POLYMER

The Bangs particles were selected for more detailed study. Their estimated

hydrodynamic radius ranges from 1030 A (Vv measurement) to 1 1 0 0 A (Hv

measurement). For uniform, monodisperse, spherical particles the expected radius o f

gyration would be a factor o f (3/5 ) 1 / 2 smaller—i.e., 800-850 A. In the absence o f added

salt, Guinier plots (Figure 3.4) gave radii o f gyration (Table 3.2) that are about 10%

56

larger than expected. The extra size may be attributed to heavy, strongly scattering

magnetite clusters located near the surface; the "halo" observed in TEM (Figure 1.2)

confirms that magnetite particles are located preferentially near the surface. Added salt

in the absence o f polyelectrolyte caused the magnetic latex to aggregate, leading to

steeper slopes in the Guinier plots. The aggregation was also evident from polarized

and depolarized DLS. The radii from Equation 3.1, 3.8, and 3.9 under various salt and

polyelectrolyte conditions are given in Table 3.2. The hydrodynamic values are

computed using the solvent viscosity.

Concentrating for the moment on the upper section o f the table, describing the case

where no polyelectrolyte has been added, and reading from left to right, it is observed

that all the various radii increase as salt is added.

The stability o f the magnetic latex particles in solutions that do not contain

polymer is determined by the magnetic attraction, van der Waals attraction and

electrical double layer repulsion. Magnetic latex particles have superparamagnetic

behavior; the magnetization is proportional to the external field, and none was applied.

The behavior o f superparamagnetic particles under applied magnetic fields was recently

reported by Fermigier and coworkers [3.32], Recently, this work was extended to non­

steady fields [3.33], In the present case, the particles' magnetic attraction energy is

smaller than that o f the van der Waals attraction [3.34] and sedimentation o f the M LP's

is a more serious problem than the magnetic interaction, though even this was

unimportant on the time scale o f several weeks due to the low concentration and small

particle size. The aggregation o f the M LP is understood simply as a result o f screening

the electrical repulsions.

In (I/

arbi

trar

y un

its)

57

□ No NaPSS/No Salt O No NaPSS/0.1 M Salt V NaPSS/0.1 M Salt

V

1 0 -

40 2 6 8 10

q2/1010 cm"2

Figure 3.4 Guinier plot o f the Bangs magnetic latex particles in pure w ater (square), MLP in the 0.1 M NaCl solution (circle), and MLP in the 0.1 M NaCl + polyelectrolytes solution (triangle), cNaPSS = 3.29 x 10" 3 g/ml.

58

Table 3.2 Radius o f magnetic latex particles in NaPSS, salt solution. These values are calculated from intensity measurements, Vv and H v dynamic light scattering measurements. H sv, HJV mean slope and intercept from H v measurement, respectively. Unit: A.

NaPSS/g ml-' No Salt 0.01 M NaCl 0.05 M NaCl 0.10 M NaCl0 Rg 900±80 9201100 1800±300 17001220

Rh(Vv) 1030120 1000115 1400175 94501300

Rh(Hsv) 1100±30 1170165 1320120 56350113000

Rh(H'v) 1110±50 12601125 1490170 34201560

7.016xl0'4 Rg 915±120 950195 9751100 930190

Rh(Vv) 1150125 1070115 1090120 1200125

Rh(Hsv) 13I0±70 1070135 1180160 1070110

Rh(H'v) 1130±60 1180170 12201100 1370140

3.292xl0'3 Rg 910±90 9201120 860160 900190

Rh(Vv) 1400±35 1260110 1250125 1200135

Rh(Hsv) 1480±90 1490160 1230160 1030120

Rh(H’v) 1210±105 1250175 13201125 1410190

6.114xl0'3 Rg 860+85 8851100 950195 9301140

Rh(Vv) 1540±25 1540135 1370135 1370130

Rh(Hsv) 1640180 1680137 1500150 1240140

Rh(H'v) 13901120 1340150 1340170 14601140

Ionic Strength [Na+1]: moles/L

’3xlO*J 9.89xl0'J 4.67x10^ 8.62x10'^

59

3.3.1.3 EFFECT OF POLYELECTROLYTE

When polyelectrolyte was added to a salt-aggregated latex suspension, the

Guinier plots returned to a more gentle descent consistent with the size o f

unaggregated latex. The effect is demonstrated in Figure 3.4. It is important to

emphasize that the scattered intensity in these experiments is totally dominated by latex:

Iiatex^poiyeiectroiyte > 70. M ost H v and Vv correlation functions, as shown in Figure 3.5,

did not exhibit pronounced nonexponentiality. Figure 3.6 (a) and (b) shows the Vv and

H v measurements o f q-dependent decay rate (T) o f magnetic latex particle in different

NaPSS concentration in pure water. The average decay rates were satisfactorily linear

with q2 for both Hv and Vv correlation functions, enabling straightforward extraction o f

D t and Dr . Figure 3.7(a) and (b) show the Vv and H v measurement o f magnetic latex

particle-NaPSS solution with 0.10 M salt concentration at various polymer

concentrations. Added salt in the absence o f polyelectrolyte caused the magnetic latex

to aggregate, leading to steeper slopes in the Guinier plots. The aggregation was also

evident from polarized and depolarized DLS.

Figures 3.8 and 3.9 show the effect o f polyelectrolyte in a convenient form. In

solutions at 0.10 and 0.05 M NaCl the translational diffusion (Figure 3.8) o f the MLP

first increases with added NaPSS and then decreases. At lower salt contents, D t

decreases monotonically. This behavior is repeated for the rotational diffusion (Figure

3.9) except that the initial increase with added NaPSS is also seen for 0.01 M salt; only

at zero added salt does a monotonic decrease occur.

Following the initial increase in diffusion rate, the mobility o f the M LP is

reduced by the continued addition o f NaPSS. The SLS data show a particle radius o f

gyration consistent with unaggregated MLP. Thus, the NaPSS polymer matrix first

stabilizes the MLP in salt-containing solutions against aggregation and then slows their

motions.

In \d

l)]

60

0

Vv

2Hv

4 ■— 0.0 0.5 1.0

t /10"2 sec

Figure 3.5 Normalized autocorrelation function vs. lag time for Bangs magnetic latex particles/NaPSS/0.1 M NaCl from Vv and H v geometries. M easurements have been done at 60° scattering angle and at 25°C .

r^./

Hz

^

61

2500- Og/ml NaPSS 7.016x10^1111 3.292xl0'3 g/ml 6.114xl0‘3 g/ml

2000 -

1500-

1000 -

500-

1064 80 2

q2/1010 cm'2

30000 g/ml NaPSS 7.016x1 O' 4 g/ml 3.292x1 O’ 3 g/ml 6.114x1 O' 3 g/ml

2500-

2 0 0 0 -

1500-

1000 -

500-

104 6 80

q2/1010 cm’2

Figure 3.6(a) Vv and (b) H v measurements o f q-dependent decay rate o f magnetic latex particles in different NaPSS concentrations in pure water.

62

3000

Og/ml NaPSS 7.016X10-4 g/ml 3.292xl0'3 g/ml 6.114x1 O’ 3 g/ml

2500-

2000 -

I '(_> 1500-

1000 -

500-

0 2 4 6 8 10

q2/1010 cm*2

2500-

2000 -

1000 -

500-

0 6 82 4 10

q2/1010 cm'2

Figure 3.7(a) Vv and (b) H v measurements o f q-dependent decay rate o f magnetic latex particles in different NaPSS concentration in 0.10 M salt solution.

63

25-

20 -

cn

1=0

ON1o 15>>CL,Cu* 10

5-

0

8Typical Error Bar

L-I *

•O.

.......

A :

"""■•'■a

••••□••■No N aC l

••-O---0.01 M N aC l

•A - 0.05 M N a C l

V • • V' • ■ 0 .10 M N aC l

0 2 4 6

^ N a P S S ^ S m l

Figure 3.8 Vv measurement o f diffusion coefficient vs. NaPSS concentration by varying NaCl condition.

64

150

100—< i

>Eexc ied

Q 50

0

Figure 3.9 H v measurement o f diffusion coefficient vs. NaPSS concentration by varying NaCl condition.

□ .1 Typical Error Bar

p...A . Q

o

V

A

□V

A V

N o NaCl

O 0.01 M N aC l

V A 0.05 M NaCl

v 0.10 M N aC l_ i ----------------------------,----------------------------(----------------------------1-----------1-----------1----------------------------r

0 2 4 6

^ N aP S S^ ^

65

For either translational or rotational motion, the retardation with added NaPSS is

greatest at no added salt. Translational motion is reduced by about 25% while

rotational diffusion is decreased by about 50% as cNaPSS is increased from 0 to

6 mg/ml. Thus, rotational diffusion appears to be more sensitive than translational to

increases in the matrix concentration at zero added salt—a surprising result which is

admittedly, established only over a small range o f cNaPSS.

Another interesting feature is the salt dependence o f the motions at a fixed,

nonzero NaPSS concentrations. Figure 3.10 shows the results. At the tw o highest

NaPSS concentrations (up and down triangles in the figure), Dt increases slightly with

added salt while Dr decreases overall. We tentatively interpret this behavior in terms o f

the ionic atmosphere surrounding the NaPSS and its effect on how closely the MLP

can approach the like-charged polyelectrolyte chains. As salt is added, the counterion

atmosphere surrounding the NaPSS shrinks. The expression given by Odijk [3.35] and

modified by Roitman [3.36], which is strictly intended for a charged cylinder, will be

used with the understanding that only the approximate significance o f ion atmosphere

effects is sought. The effective electrostatic diameter surrounding the chain, deg-, is

predicted to be:

Here b is the distance o f charge separation along the chain contour, d0 is the hard core

where N a is Avogadro's number, I is the ionic strength o f the solution, and B e is the

Bjerrum length,

(3.11)

diameter o f the chain, which deff will approach in the limit o f high salt, k ' 1 is the

Debye-Huckel electrostatic screening length given by

k ‘ 1 = (STrNaiBg/lOOO) ' ^ 2 (in centimeters) (3.12)

ekT(3.13)

Vv,

D/1

0'8 c

m2s

'

66

□ 0 g/ml NaPSS O 7.016X10"4 g/ml A 3.292xl0'3 g/ml

- 2 -

0.120.080.00 0.04

£

C/3

NaCl Cone. [M]

Figure 3.10 Translational (left axis) and rotational diffusion (right axis) o f the MLP varying salt concentration.

67

where e is the charge o f an electron, 4.8 x 10' 1 0 esu, and e is the dielectric constant o f

the medium, 78 for water. This leads to Be = 7 A at T = 298 K. A reasonable interion

distance is b = 2.5 A [3.37] and the core diameter o f d0 = 1 2 A can be obtained,

approximately, from p 7ibd e f f 2 / 4 = M</N a where p is the polymer density (assuming p =

1 g/mL is adequate for the present purpose, as d0 is not a critical parameter) Mj, is the

monomer molecular weight, 183 g/mol for the fully ionized monomer. The effective

diameter increases from about 60 A to about 2 0 0 A as salt is decreased from 0 . 1 M to

0.01 M.

How significant is this increase, in terms o f diffusion o f the magnetic latex

probes? Many models have been devised to describe the reduction o f translational

diffusion due to space occupied by a polymeric matrix; for a sampling, see ref. [3.38],

A representative expression is Dt = D°t(l-(|)2 ) where <j)2 is the polymer volume fraction

and D°t is the translational diffusion in the simple solvent (4>2 = 0). In the case o f a

polyelectrolyte retarding the diffusion o f a like-charged probe, the polymer volume

fraction can be estimated as (j) 2 « (7ibd2 efI/ 4 )(c/M 0 )Na. When deff = d0 = 1 2 A this gives

<J)2 = c x 1 mL/g (d0 was effectively chosen to make this true). The five-fold increase to

deff = 60 A at 0 . 1 M NaCl causes a 25-fold increase in the effective value o f <}>2 while

the 16-fold increase to deff = 2 0 0 A leads to a 256-fold increase. At the maximum

concentration o f 0.006 g/mL, one finds <J>2 = 0.15 and 1.5 respectively. The latter value

is unphysical due to the many assumptions o f this oversimplified model. Nevertheless,

significant changes in D t due to changes in the effective electrostatic diameter o f the

polyelectrolyte should be expected. The effect as far as rotational diffusion goes

appears to be that the polyelectrolyte can approach the MLP more closely as salt is

added and therefore retard its rotational diffusion more effectively. Previous results on

the self diffusion o f negatively charged bovine serum albumen in the presence o f

68

negatively charged DNA also showed an increase in translational diffusion with added

salt [3.39], Rotational diffusion data were not obtained.

W e included the above section on ion atmosphere effects because o f the

difference in behavior for rotational and translational diffusion. The former is clearly a

decreasing function o f salt, while the latter depends weakly on salt, but increases

slightly at the highest cNapSS. Increasing translational diffusion with added salt could as

easily be explained by changes in the solution viscosity. Indeed, viscosity

measurements show that the increase in Dt with added salt would be larger than it is, if

viscosity were the controlling factor. The information provided by the rotational

diffusion behavior compels us to consider the ion atmosphere effects, which can explain

the decreasing rotational mobility as the result o f closer approach to the NaPSS chains.

The present work is ill-suited to test the validity o f the Stokes-Einstein relation

due to the limited range o fN aPSS concentrations. According to Equation 3.5 and 3,7,

the products Dtr| and Drr| are constants in simple fluids. Here we will comment only

that these products are not constants for M LP in the relatively dilute solution ofN aPSS

studies. M odest increases (ca. 15%) o fD tri and Drr| were observed as cNaPSS was

varied.

3.3.2 FLUORESCENCE PlIOTOBLEACHING RECOVERY

3.3.2.1 LABELING THE POLYSTYRENESULFONATE SODIUM SALT

To test for binding o fN aPSS to MLP, fluorescently tagged NaPSS was

prepared for optical tracer diffusion measurements by the fluorescence photobleaching

recovery method. Labelling had been done by Mr. A. Davila in Dr. M. L. Mclaughlin's

group in the chemistry department here at LSU. The attachment o f fluorescent dye to

NaPSS was a tw o step process [3.40], In the first step the sodium sulfonate was

substituted with the sulfonyl chloride. Next, the sulfonylchloride was reacted with an

amine containing fluorophore, fluoresceinamine isomer I (Aldrich) (Figure 3.11).

69

+ POCI3

(excess)

CH CH2"

;ojS02C1

(a)

NH

CH CH2"

s o 2 c i

+ COOH

HO

THFEt-iN

(excess)

LPSSCL1) NaOH, H20

----------------12) HC1,H20

CH— CH2 -

CHS02NHDye

- CH— CH2

SQ3Na0 . 0 0 2

(b)

NaPC^Clj

LPSSCI

0.998

Figure 3.11 Mechanism o f the labeling the NaPSS with fluorescent dye. (a) First step is to substitute the sodium sulfonate with the sulfonyl chloride and (b) then the sulfonylchloride is reacted with amine containing fluorophore.

70

Three grams ofN aPSS was suspended in 30 ml o f phosphorous oxychloride,

POCI3 , in a 100 ml round bottom flask. The mixture was stirred and refluxed for 2

days, then cooled to room temperature. W ater was added to neutralize the remaining

POCI3 . The resulting solid was filtered and washed with H20 until neutral pH. The

product was placed in an oven at 80°C over 3 days, after which 1.90 g o f a beige solid

was obtained. Decomposition temperature range o f both the starting material and the

product were checked during the each step (starting material 160 (white)-200°C

(white), product 160 (brown)-200°C (black)). In a 100 ml Schlenk flask, 1.16 g o f the

product o f the first step was suspended in 20 ml o f dry tetrahydrofiiran, THF. The

polymer suspension was stirred at room temperature. Fluoresceinamine (0.01 g,

0.03 mmol) and triethyamine (0.87 g, 8.64 mmol) were dissolved in 5 ml o f dry THF.

The dye-THF solution was heated to boiling to ensure complete dissolution, then added

dropwise to the product o f the first step. Color changes signaled instantaneous

reaction. The THF was removed at reduced pressure. The yellow residue (labeled

poly(styrenesulfonylchloride), LPSSC1) was dissolved in 10 ml o f 10% sodium

hydroxide and boiled for 20 minutes to hydrolyze unreacted sulfonyl chlorides. Then

3 M HC1 was added until the pH was neutral. The resulting bright yellow gel was dried

in an oven at 80°C for 24 h.

3.3.2.2 CHARACTERIZATION OF LABELED PSS (LPSS)

Even after filtering the final product o f labeled NaPSS with 0.1 pm Millipore

VV filter, the hydrodynamic radius o f the labeled NaPSS (LPSS) was bigger than that

o f the starting materials (NaPSS). The diffusion o f the NaPSS was S Jx lO ^ cm ^ /s ' 1

and the labeled NaPSS was 2.6x1 O' 7 cm'2 /s_1. The diffusion coefficient from DLS at

25°C o f the NaPSS was 2 .4x l0 ' 7 cm ^/s ' 1 under conditions o f high added salt (0.1 M).

The difference is essentially insignificant. Figure 3.12 shows the DLS comparison

between the LPSS and starting NaPSS in different salt condition.

71

_After dialized and freeze dried (LPSS)1.0

Just after synthesis (LPSS)0.8

0.6

Q.|̂ 0.4

0.2

0.0

0.01 1000 100000.001 0.1 10 100

Gamma (ms)

A fte r a d d in g sa lt (0 .5 M )1.0

NaPSS

0.8

LPSS

0.6

O.

^ 0.4

0.2

0.0

1000 1000010 1000.001 0.01 0.1 1

Gamma (ms)

Figure 3.12 DLS comparison between LPSS and NaPSS in different salt conditions. At high salt condition, hydrodynamic radii o f LPSS and NaPSS are the same.

72

The visible absorption maximum o f the labeled polymer (497 nm) was close to that o f

fluorescein salt in water (491 nm). On the assumption that the absorption coefficient

was similarly unaffected, we estimate the efficiency o f dye labeling as 0.5 dyes attached

per polymer chain, corresponding to 0.24% o f monomer subunits labeled assuming a

labeled molecular weight o f 70,000. It is assumed that this low level o f dye will not

perturb the fairly structureless NaPSS random coils.

3.3.2.3 FLUORESCENCE PHOTOBLEACHING RECOVERY MEASUREMENT

The FPR instrument, conceptually similar to that o f Lanni and W are [3.41] has

been described recently [3.42], During measurement, an intense laser beam briefly

illuminates a coarse diffraction grating placed in the rear focal plane o f an

epifluorescence microscope to generate, by photobleaching, a square wave pattern o f

spacing L in the fluorescent sample. Then the grating is translated at a constant speed

and projected into the sample by a less intense beam. As its image falls into and out o f

coincidence with the bleached pattern, a weak ac signal is produced, on top o f a large

dc baseline from the unbleached fluorophores, at the anode o f a photomultiplier tube

that monitors the illuminated region. Initially, the ac signal resembles a triangle wave.

As bleached and unbleached diffusers move about, the contrast o f the bleached pattern

is reduced. The ac signal loses its sharp edges and approaches a sine wave. The wave

amplitude o f the fundamental component decays at a rate, T f, proportional to the

diffusion coefficient o f the fluorescent diffusers [3.41]: ac( t) = ac(0)exp(-rjt), with K

= 27i/L and T = DK2. The spatial frequency K may be varied by changing the coarse

diffraction grating or the microscope objective. Bleach depth (the percentage o f dyes

photobleached) was less than 25% for all the measurements, and this level o f

photobleaching was reached with exposure times less than 0 .1 5 P 1 The present

conditions will be adequate for our purposes. The decay rate T is obtained by fitting

73

the wave amplitude o f the fundamental to a single exponential with floating baseline,

using a M arquardt nonlinear least squares algorithm. LNaPSS was measured in water

at several K values, and T varied linearly with K2, indicating that chemical recovery o f

the dye and/or convection were insignificant. M ost experiments with LNaPSS in MLP

solutions were conducted at just one K value. All FPR measurements were at 25.0 ±

0.2°C.

3.3.2.4 STABILIZATION MECHANISM

The most intriguing aspect o f the combined SLS and DLS results is the

stabilization o f M LP against salt-induced aggregation that is effected by NaPSS. This

result is independent o f the order o f mixing: NaPSS will prevent salt-induced

aggregation o f M LP—or undo it if it has already occurred. It seems likely that the MLP

may have adsorbed surfactants or oligomers o f unknown composition in order to

prevent strong binding in the flocculated state. W e are aware o f just one other report

where dilute polymer caused salt-flocculated colloids (deliberately stabilized against

very tight binding) to be resuspended [3.43], In the present case (and probably also in

Ref. [3.43]) it is expected that this stabilization effect is due to free polymer. The

reasoning in the present case is that both M LP and NaPSS are negatively charged. It is

possible that poorly sulfonated zones could bind to weakly charged patches on the latex

particles o r displace some previously attached stabilizer, leading to a sort o f steric

stabilization. However, elemental analysis suggests that such poorly sulfonated zones

are virtually absent from this sample ofN aPSS. Also, the magnetic latex particles used

in this experiment have high charge density and tight parking area (« 3A2/COOH

group). Still, it was decided to test for possible binding by monitoring the diffusion o f

fluorescently labeled NaPSS (LNaPSS) in the presence o f the latex spheres using FPR.

74

Any LNaPSS that is permanently bound to the purified Bangs M LP would

diffuse some 10 times more slowly than free LNaPSS. In this case, the fluorescence

recovery would be characterized by tw o exponential signals:

A C (t) = xf exp"K2Dft+ xb exp-KDb‘ (3.14)

where AC(t) is the magnitude o f the ac signal produced by the modulation detector

(see experimental section) and xf and xb are mole fractions, and D f and D b the diffusion

coefficients, o f LNaPSS that are free and bound, respectively. It is assumed that the

fluorescence intensity and photobleaching efficiency are not modified by the binding

process. I f the LNaPSS undergoes rapid exchange with the M LP on the time scale o f

the FPR measurement, several tens o f seconds as performed here, then a single

exponential recovery profile will be observed, and an average diffusion coefficient will

be obtained: D avg = xbD b + X(Df.

Figure 3.13 shows an FPR trace from a solution containing LNaPSS and Bangs

MLP. The semilogarithmic representation demonstrates that the trace is effectively

single exponential. The inset shows a linear increase in recovery rate with K 2 and zero

intercept, characteristic o f pure diffiisional relaxation o f the pattern photobleaching.

The average diffusion coefficient is effectively that o f the free LNaPSS and this remains

true even at very low concentrations o f LNaPSS and high concentrations o f MLP.

Figure 3.14 shows the diffusion coefficient o f LNaPSS as a function o f the quotient o f

total particle surface area and the area required to bind all LNaPSS molecules to a

surface. The former area (per unit volume o f solution) is simply vMLP(4 7 rR2MLP),

where symbols v and R represent the number density and hydrodynamic radius,

respectively. The number density was computed using a density o f 2 g/ml for the M LP

(the actual density o f the unrefined Bangs particles is 2.27 g/ml).

Ln (P

eak

AC

-Bas

eline

) Pe

ak

AC

/V

75

2

1

0

100 125500 25 75

t/s

- 2 -

■H* ++++-4-

- 6 -

100 125

Figure 3.13 FPR traces for LNaPSS (cLNapss = 15 x 10' 5 g/ml) in 1.8 wt% MLP suspension. Top: exponential decay and single exponential fitted curve. Inset: decay rate scales linearly with K2. Bottom: ln (peakAC-baseline) vs. time plot.

76

The area required to bind all the polymer was computed as vLNapSS(4 7 tR 2 LNapSS) and

the relation vLNaPSS = N a(cLNapSS/M w) where N a is Avogadro's number and M w is the

weight average polymer molecular weight, 70,000. Thus, the area covered by an

LNaPSS molecule is the projection o f the expanded LNaPSS chain in solution onto a

surface. It is understood that the actual area covered per chain may differ due to

conformational changes associated with binding and/or interpenetration o f chains.

Also, our simple model does not take into consideration polydispersity o f the LNaPSS,

surface roughness o f the M LP due to near-surface magnetite inclusions, or possible

curvature effects for the two MLP sizes used to generate Figure 3.14 (unrefined

0.8 pm Bangs latex spheres were used at the highest concentrations). In brief, the area

quotient o f Figure 3.14 is not absolute, but represents a fixed set o f assumptions.

However, those assumptions do give to surface-bound LNaPSS a generous area— and

still there is, at the highest M LP concentrations, more than enough area at the surface

o f the MLP to accommodate all LNaPSS chains. Yet there is no decrease in the

average diffusion coefficient to signal that any chains are bound, and no

nonexponentiality to indicate permanent binding. Specific binding o f LNaPSS to MLP,

either permanent or by rapid exchange, seems unlikely.

3.4 SUMMARY

The DLS and SLS results from the M LP/NaPSS system show : 1) translational

and rotational diffusion coefficients o f magnetic latex particles are decreased by

increasing the matrix concentration at the no salt condition; 2 ) the diffusion o f magnetic

latex particles in the no NaPSS condition is abruptly dropped by increasing the salt

concentration to 0.1 M, and static light scattering data shows aggregation o f the

particles in high salt conditions when there is no NaPSS; 3) aggregation o f the particles

can not be observed and diffusion increases with increasing NaPSS concentration in

high ionic strength systems.

77

cs's0r-1o

£ r

8

6

4

2

0.0000 0.001 0.10.01 1 10Total Surface Area of MLP/Possible Occupied Area of NaPSS

Figure 3.14 Self diffusion o f LNaPSS as a function o f the quotient o f M LP surface area and the area required to bind all LNaPSS (see text). 0.2 pm M LP was used for this experiment except the last tw o (full) points for which 0.8 pm M LP was used. Added salt: 0 . 1 M.

78

These inter-particle interactions are caused by the der Waals attraction and the

electrical repulsion.

The polymer stabilization o f the particle flocculation in this experiment may be

governed by the electrically free polymer rather than the steric mechanism. Based on

the self diffusion coefficient o f the FPR measurement, the depletion stabilization o f the

M LP in the polyelectrolyte matrix has been proposed.

79

3.5 REFERENCES

3.1 (a) L. B. Bangs, "Uniform Latex Particles," Seradyn Inc., Indianapolis, IN 1984.

(b) General Booklet o f Rhone-Poulenc, "Estapor," 1987.

3.2 (a) M. Fermigier and A. P. Gast, J. Coll. Inter. Sci. 154, 552 (1992).

(b) M. Fermigier and A. P. Gast, J. Magnetism and Magnetic Materials 122, 46 (1993).

(c) D. W irtz and M. Fermigier, Phys. Rev. Lett. 72, 2294 (1994).

3.3 (a) G. D. J. Phillies, J. Phys. Chem. 93, 5029 (1989).

(b) P. S. Russo, M. Mustafa, T. Cao, and L. K. Stephens, J. Coll. Inter. Sci. 120,122(1988).

(c) S. Gorti and B. R. Ware, J. Chem. Phys. 83, 6449, (1985).

3.4 R. Piazza, J. Stavans, T. Bellini and V. Degiorgio, Optics Comm. 73, 263 (1989).

3.5 (a) M. M ustafa and P. S. Russo, J. Coll. Inter. Sci. 122, 120 (1988).

(b) T-H. Lin and G. D. J. Phillies, Macromolecules 17, 1686 (1984).

(c) W. Brown and R. Rymden, Macromolecules 19, 2942 (1986).

3.6 D. Sohn, M. L. DeLong, and P. S. Russo, Mat. Res. Soc. Sym. Proc. 248, 247(1992).

3.7 J. C. Brown, P. N. Pusey, J. W. Goodwin, and R. H. Ottewill, J. Phys. A: Math. Gen. 8(5), 664 (1975).

3.8 R. Krause, G. Nagele, D. Karren, J. Schneider, R. Klein, and R. W eber, PhysicaA 153, 400 (1988).

3.9 E. J. W. Verwey and J. T. G. Overbeek, "Theory o f the Stability o f Lyophobic Colloids," Elservier Publishing, NY, 1948.

3 .10 D. H. Napper, "Polymer Stabilization o f Colloidal Dispersion, Academic Press," London, 1983.

3.11 J. E. Johnson and E. Matijevic, Coll. Polym. Sci. 270, 364 (1992).

3.12 M. Sedlak and E. J. Amis, J. Chem. Phys. 96(1), 826 (1992)

3.13 J. M eadows, P. A. Williams, M. J. Garvey, and R. Harrop, J. Coll. Inter. Sci. 148(1), 160(1992).

80

3.14 K. Furusawa, Z. Shou, and N. Nagahashi, Coll. Polym. Sci. 270, 212 (1992).

3.15 H. Inoue, H. Fukke, and M. Katsumoto, J. Coll. Inter. Sci. 148(2), 533 (1992).

3.16 (a) S. Asakura and F. Oosawa, J. Chem. Phys. 22, 1255 (1954).

(b) S. Asakura and F. Oosawa, J. Poly. Sci. 33, 183 (1958).

(c) A. P. Gast, C. K. Hall, W. B. Russel, J. Coll. Inter. Sci. 96, 251 (1983).

3.17 R. I. Feigin and D. H. Napper, J. Coll. Inter. Sci. 75, 525, (1980).

3.18 (a) F. Oosawa, "Polyelectrolytes." Marcel Dekker Inc., N .Y ., 1971.

(b) N. A. Rotstein and T. P. Lodge, Alacromolecules 25, 1316 (1992).

(c) J. Newman and K. L. Schick, Biopolymer 28, 1969 (1989).

(d) R. Bansil, S. Pajevic, and C. Konak, Macromolecules 23, 3380 (1990).

3.19 (a) R. S. Koene and M. Mandel, Macromolecules 16, 220 (1983).

(b) R. S. Koene, H. W. J. Smit, and M. Mandel, Chem. Phys. Letters 74(1), 176 (1980).

3.20 (a) M. Drifford and J. -P. Dalbiez, Biopolymers 24, 1501 (1985).

(b) M. Drifford and J. -P. Dalbiez, J. Phys. Chem. 8 8 , 5368 (1984).

3.21 (a) W. B. Russel, "The Dynamics o f Colloidal Systems," Univ. o f WisconsinPress, WI, 1987.

(b) T. Sato and R. Ruch, "Stabilization o f Colloidal Dis persions by Polymer Adsorption," Marcel Dekker, Inc., N.Y., 1980.

3.22 R. Buscall, J. Chem. Soc. Faraday Trans. 1, 77, 909 (1981).

3.23 (a) A. Guinier and G. Foum et, "Small-Angle Scattering o f X-rays," John Wileyand Sons, NewYork, 1955.

(b) M. Kerker, "The Scattering o f Light and Other Electromagnetic Radiation," Academic Press, N ew York, Ch. 8 , 1969.

3.24 B. Berne and R. Pecora, "Dynamic Light Scattering." Wiley, N ew York, 1976.

3.25 P. S. Russo, M. J. Saunders, L. M. DeLong, S. K. Kuehl, K. H. Langley, and R. W. Detenbeck, Anal. Chem. Acta 189, 69 (1986).

3.26 C. Tanford, "Physical Chemistry o f Macromolecules," Wiley, N ew York, 323 (1961).

81

3.27 (a) T. -L. Lin and G. D. J. Phillies, Macromolecules 17, 1686 (1984).

(b) G. S. Ullmann, K. Ullmann, R. M. Lindner, and G. D. J. Phillies, J. Phys. Chem. 89, 692 (1985).

(c) P. S. Russo, L. K. Stephens, T. Cao, and M. Mustafa, J. Coll. Int. Sci. 122, 120 (1988).

3.28 B. Bem e and R. Pecora, "Dynamic Light Scattering." Wiley. New York, 1976.

3.29 D. E. Koppel., J. Chem. Phys. 57, 4814 (1972).

3.30 J. -C, Daniel, J. -L, Schuppiser, and M. Tricot, U. S. Patent, Nov. 9, 4,358,388 (1982).

3.31 Personal communication with S. R. S. Aragon at San Francisco State University.

3.32 a) M. Fermigier and A. P. Gast, J. Coll. Inter. Sci. 154, 552 (1992).

b) M. Fermigier and A. P. Gast, J. Magnetism and Magnetic Materials 122, 46 (1993).

3.33 D. W irtz and M. Fermigier, Phys Rev. Lett. 72, 2294 (1994).

3.34 R. E. Rosensweig, "Ferrohydrodynamics," Cambridge, N.Y., 49 (1985).

3.35 T. Odijk, Macromolecules 19, 2313 (1986).

3.36 D. B. Roitman, R. A. Wessling, and J. McAlister, Macromolecules 26, 5174(1993).

3.37 N. Borochov and H. Eisenberg, Macromolecules 27, 1440 (1994).

3.38 S. Pickup and F. D. Blum, Macromolecules 22, 3961 (1989).

3.39 M. R. W attenbarger, V. A. Bloomfield, Z. Bu, and P. S. Russo, Macromolecules 25, 5263 (1992).

3.40 Organic Synthesis I. 84.

3.41 F. Lanni and B. R. Ware, Rev. Sci. Instrum. 53(6), 905 (1982).

3.42 Z. Bu and P. S. Russo, Macromolecules 27, 1187 (1994).

3.43 J. E. Seebergh and J. C. Berg, Langmuir 10, 454 (1994).

CHAPTER 4

KINETIC STUDIES OF MAGNETIC LATEX PARTICLES' SELF ASSEMBLY UNDER APPLIED MAGNETIC FIELD

82

83

4.1 INTRODUCTION

The formation o f large clusters by the union o f many separate, small elements

includes a wide variety o f processes in polymer science, in colloidal physics, and in

materials science. These aggregation processes have been studied for a long time with

computer simulation using scaling and fractal concepts, or experimentally with optical

microscopy and scattering techniques. Kinetic studies o f these aggregation processes

mainly involve the determination o f the cluster sizes on a certain time scale, s(t), and

the cluster size distribution function, N(t). Based on the theoretical and experimental

methods, diffusion limited (DLA) and reaction limited (RLA) aggregation models have

been proposed [4.1]. Gelation and colloid aggregation are examples o f spontaneous

aggregation processes. An induced electric or magnetic field also changes the kinetics

o f the cluster growth.

Electric field induced kinetic studies have been performed mainly on

electrorheological (ER) fluids, which consist o f a suspension o f fine dielectric particles

in a liquid o f low dielectric constant and low viscosity [4.2], In a large electric field, a

few kV/mm, polarized particles first form chains along the field lines to form a one­

dimensional solid, and these chains slowly aggregate to form columns. The apparent

viscosity o f the fluid increases dramatically in the presence o f an applied electric field.

These phenomena were first reported by Winslow in 1949, and so are referred to as the

"Winslow effect." In the past decade many industrial applications have been

implemented such as variable transmissions, shock absorbers, variable pumps, and so

on [4.3],

The dipole interactions o f polarizable and magnetizable particles are basically

identical when the particles are uniform spheres. In the presence o f an electric field,

induced dipole moment changes proportionally with the strength o f external field E

[4.4]

84

M- = e wRE p E w

ep + 2 ew(4.1)

where e^ ,, 8 p are the complex dielectric constants o f the solvent and the particle,

respectively, R is the radius o f the sphere, and E is the electric field.

In the case o f a magnetic field,

p = p 0 - j 7 d l V (4.2)

where % is the magnetic susceptibility, p 0 is the magnetic permeability o f the vacuum,

and H is the external field. For some magnetic fluids, for example ferrofluids, it has

been found that large agglomerates o f particles are formed in the presence o f a weak

magnetic field. The needle-like agglomerates appear when a field o f several Gauss is

applied. The needle-like structures disappear when the magnetic field is removed. This

is because o f the paramagnetism o f the particles. Many investigations o f aggregation in

ferro-fluids have been carried out using experimental and computational methods [4.5],

Two basic considerations o f the magnetic field studies and a general approach to the

cluster aggregation are dealt with in the next section.

4.2 KINETIC MECHANISM AND THEORY

4.2.1 M E C H A N IC A L C O N C EPTS

D e Gennes and Pincus applied mechanical considerations to identical spherical

ferromagnetic grains o f radius R (diameter d) suspended in a magnetically passive

liquid [4.6], W hen the particles do not carry any electrostatic charge, the dipole-dipole

interaction at distance a is defined through a dimensionless constant X. Neglecting

higher order interactions, tw o spheres with aligned identical dipole moments |i interact

through a potential energy at closest distance,

(j. = M(rca3 / 6 ), where M is magnetization, 0 is the angle between dipole moment and

center-to-center vector, and X is proportional to the volume per grain. From Equation

4.2 and 4.3,

where the magnetic susceptibility, x, can be determined by a magnetophoresis

experiment or data o f Bossis et al. [4.7], The mean number o f particles, s(t), in an

equilibrium chain depends on the dipole strength.

These concepts and theoretical considerations were extended by Jodan [4.8] and

Krueger [4.9], and their predictions have been supported by several experiments such

as light scattering [4.10], microscopy [4.11], and pulsed magnetization [4.12],

4.2.2 PH A SE TR A N SITIO N C O N C EPTS

K. Sano and M. Doi explained these phenomena with the concept o f the gas-

liquid phase transition, the gas phase o f magnetic particles is transformed to the liquid

phase by the magnetic field [4.13], They consider a) the dipole-dipole interaction (X'),

b) the dipole-field interaction (h = p/Z/kT), and c) the van der W aals interaction (e') and

discuss the phase boundary with following equation,

where oc = V/N is the volume per particle in the most closely packed state; v is the

coordination number; c = Np/N where N p is the number o f the particle in the system; m

= M tot/Npp, m = 1 is the state o f the magnetic moment if the particles are completely

aligned; N p is the number o f the particles in the system; M tot is the total magnetization;

^ U (a = 2R ,6 ~ 0) ftp 0 R V #3 2 r j 2

kT 9kT(4.4)

2 9

s ( t)= 1 - — 4>X exp(2X)-l

(4.5)

(4.6)

86

X is the dipole-dipole interaction parameter; and e is the van der Waals' contribution.

The phase separation o f the ferromagnetic fluid is described by the dipole-dipole

interaction (X'), and van der W aals interaction (s'). They predict that phase separation

occurs for finite magnetic field strength at very large values o f X.

4.2.3 KINETIC CONSIDERATIONS

In the DLCA model, clusters stick to each other irreversibly and immediately,

and the kinetics is limited by only the diffusion o f the cluster. The fractal dimension, df,

in this case is ~ 1.75±0.05. In the RLCA case, the sticking probability is much less than

unity, so many collisions will occur before two clusters link together. These dynamics

are much slower and df ~2.05±0.05. Fractal geometry is a natural description for

disordered objects ranging from macromolecules to the earth's surface [4.14],

Experimentally, colloid particles like gold clusters [4.15], sol-gel transitions with silica

aerogel [4.16], and magnetic particles like ferromagnetic particles have been considered

for the study o f tw o or three dimensional fractal structure. These kinetic theories are

extended to monomer-cluster and cluster-cluster processes.

The kinetics o f field induced cluster aggregation processes is based on the

diffusion limited model by M eakin [4.17] and Kolb et al. [4.18]. These m ethods are

developed for the cluster size distribution function, s(t), using the number o f cluster,

N(t), and size s at time t.

s(t) = Z N ( t ) s 2 octz (4.7)£ N ( t ) s

Two-dimensional simulation results show the mean cluster size, s(t), scales according

to time with the exponent z. For the actual fitting in the experiment, we use the

following equation.

s(t) = s(0)[l + ( t / T ) z ] (4.8)

87

The characteristic time or growth rate x describes the response o f the sample to the

magnetic field given by the rate o f cluster growth per unit length. It depends on the

strength o f the magnetic field and concentration o f the sample.

Superparamagnetic particles behave like normal latex in solution with

electrostatic repulsion and van der Waal's attraction competing in the usual way. When

a magnetic field is applied to these particles, they align along the field direction and the

cluster size grows in time. A simple model for such an aggregation for oriented

paticles has been proposed by Miyazima and coworks [4.19]. Miyazima et al. provide

the scaling theory o f the diffusion-limited aggregation which describes the time

evolution o f the cluster size distribution. Their aggregation process o f oriented

anisotropic particles indicates that, asymptotically, s(t) ~ tz and that the exponent z is

1 / ( 1 -y) where y is the exponent describing the dependence o f the diffusion coefficient

on the cluster mass. They assume that the diffusion coefficient D , for clusters o f size s

is given by D s = D°sY where D° is a constant and y is the exponent describing the

dependence o f the diffusion coefficient on the cluster mass. Fraden et al. determined z

= 0.6 ± 0.02 and y = -0.67 with 1.27 pm dielectric spheres at interaction strength X =

U/kT = 30.9. Fermigier et al. reported that mean cluster size increases approximately

as t0 5 with 1.5 pm paramagnetic particles and assuming y = -1. Janssen et al. [4.20]

reported a 4/3 exponent for the initial coagulation rates o f Mn2 C> 3 particles.

The time dependent size distribution o f the magnetic cluster is investigated here

by video microscopy and video small angle light scattering. The length scale o f the

clusters from these two methods is considered and the scaling constant is also

determined from video microscopy.

88

4.3 VIDEO MICROSCOPY

4.3.1 GENERAL INTRODUCTION

Optical microscopy is an old technique, but it is an extremely useful one having

a variety o f applications. There are many different kinds o f optical microscopy

techniques such as polarized light microscopy [4.21], fluorescence microscopy [4.22],

IR, U V or Raman microscopy [4.23], laser and holographic microscopy [4.24],

interface microscopy [4.25], and phase contrast microscopy [4.26]. Video microscopy

is a simple combination o f a television camera, a video monitor, a light microscope, and

a computer. There are advantages in using a combination o f video and optical

microscopy, the main one being ease with which imaging data can be handled. This is

especially true for kinetic studies. Video techniques can also substantially increase the

image contrast by adjusting the black or threshold level o f the video signal [4.27],

Another benefit is the possibility o f imaging at very low light levels, which is possible

by using a silicon intensified target tube or solid state camera [4.28], These techniques

are commonly applied to kinetic studies in biological laboratories. In this chapter, the

evolution o f structures based on magnetic latex particles under an applied magnetic

field has been investigated with the video microscopy technique.

4.3.2 CELL GEOMETRY AND MAGNETIC FIELD

The cell geometry with the applied magnetic field and the respective scales are

shown in Figure 4.1. M icro slides manufactured by Vitro Dynamics Inc. were used as

sample cells. They have a 0 . 1 mm path length and rectangular shaped cross section.

This thin chamber is placed inside another 4 mm diameter glass tubing, which has 240

turns o f 26 HF #1 copper wire on each side. The magnetic field is created inside o f the

0.4 mm tubing by applying electric current to the copper wire. This Helmholtz coil

configuration has a 7 mm spacing in the middle o f the cell for microscopic observation.

89

Microscopy Observation

7 mm

240 tu m s/3 8 mm

4 mm cylindrical tubing

0.1 mm x 2 mm Vitro Cell

SALS Laser Beam

Figure 4.1 Cell geometry, Helmholtz coil, to induce the magnetic field.

90

A Heath 2718 DC tri-power supply generates up to 15 A. The magnitude o f the

homogeneous magnetic field in the middle o f a solenoid is [4.29],

where B is the magnitude o f the axial field at the midpoint on the center axis, p 0 is the

permeability o f free space, p 0 = 4 % 10" 7 T m A"1, N is the number o f turns with axes

coinciding, I is the applied electric current, rt is the radius o f the tubing, dt is the

spacing between the tubing, and x = dt/2. Geometrical difficulties o f Helmholtz coil

such as spacing and number o f turns with axes coinciding constraint from the

calculation o f actual magnetic field strength. The actual magnetic field between the

tubing is measured by the Gauss/Tesla m eter where a detecting probe, 4 mm wide, can

be positioned at the middle o f the Helmholtz coil. The actual value o f the magnetic

field strength, ~10 G, is smaller than the theoretical calculation (~ 130 G for 1 A

current). This magnetic field induced device is mounted on the microscope stage with

insulating tape.

High current heats the sample, which causes increased thermal motion o f the

particles and, eventually, solvent boiling. The thin capillaries were blown apart, so

experiments were done with applied current under 1.5 ampere.

4.3.3 SA M PL E PR E PA R A T IO N

M agnetic latex particles (MLP) were purchased from Bangs Laboratories, Inc.

They contained 6 8 % iron oxide by weight and had a mean advertised diameter o f 0.8

pm. The aqueous solution provided at 8 w t% was treated with a surfactant (sodium

dodecyl sulfate) to minimize the sedimentation and charged interactions. To obtain

more homogeneous particles, we diluted the solution ( 1 0 times) and kept it in the

cylinder for 3 days to settle the big particles, then collected aliquots from the middle o f

B _2 d,

91

the solution. The selected solution was diluted again with pure water and 0.2 wt% , 0.1

wt% , and 0.05 wt% stock solutions were prepared. Each concentration was confirmed

by thermal analysis. These samples were transferred to Vitro dynamics micro cells.

They were then placed in a Helmholtz coil stage. Because o f particle sedimentation,

measurements were made as quickly as possibly and performed by changing the sample

for each different field experiment.

The meaningful parameters o f microscopic observations are the number o f

particles N(t), solid surface fraction, mean cluster size, and characteristic time. The

effective solid surface fraction <J)e can be determined from the total initial number o f

particles N(0) in the area detected by the video camera, A (162 pm x l2 8 pm ) as [4.30]

<j>e = N (0 ) tcR2 /A (4.10)

where R is the radius o f the particle.

4>e is 0.06 for the 0.2 wt% , 0.03 for the 0.1 wt%, and 0.02 for the 0.05 wt% sample,

respectively.

4.3.4 IM A G E PR O C E SSIN G

The Nomarski condenser o f the Leitz Metallux 3 microscope permits increased

brightness and resolution. The Leitz Metallux 3 microscope is connected to the M TI

SIT (silicon intensified target) 6 6 video camera, and the microscope image is recorded

on a JVC VCR. The M acintosh Scion imaging process (or running N IH image 1000

software) is used to digitize the image. The software is then used to measure the size

o f each cluster. One can collect the digitized image according to the different time

scale, using the function "Make Movie." Each image is followed by "Threshold" and

"Reduce Noise," and the size o f each cluster is measured. This data file was saved on

the M acintosh Ilfx com puter read-write optical disk, and exported to an IBM

compatible computer. All data analysis has been performed using the data analysis

program ORIGIN.

92

IBM Computer

Printer or Video Printer

M i l Sit 66 Video Camera

Kepco or Heath Power Supply (0-20 A)

Magnetic

VCRMacintoshScion Imaging System

Monitor

Figure 4.2 Experimental set up for the video microscopy experiment.

93

4.3.5 R E SU LT S AND D ISC U SSIO N

Under a high enough magnetic field, it is reported that the mean cluster size

increases as t0 5 (Gast et al.) or t° 6 (Fraden et al.) [4.31], However, ours are weak

external fields where the aggregates can reach an equilibrium size, their attractive

dipolar interaction being balanced by thermal forces.

Figure 4.3 shows a set o f microscopic pictures o f growing M LP under the

magnetic field (0.5 A current, 6 Gauss). At the beginning, each particle undergoes

random thermal motion and then begins to align with others due to the applied field.

The cluster size increases and then reaches a saturation point. Figure 4.4 and 4.5 show

the measured cluster size versus applied magnetic field with different concentrations.

The initial cluster size is not exactly the same as the real particle size because the

background image effects are more serious at small time scales. The initial point o f the

plot is fitted to the original particle size, 0 . 8 pm and couple o f initial points are

calibrated with the background image.

By increasing the magnetic field and sample concentration, the rate o f cluster

growth is faster. The rate o f the cluster growth, x '1, can be calculated from the kind o f

data shown in Figures 4.4 and 4.5. From the slope and intercept o f the ln(s(t)/s(0)-l)

vs. ln(t) plot (Figure 4.6) z and x ' 1 are derived and summarized in Table 4.1.

According to Equation 4.8 x is a constant for each different magnetic field. An average

exponent o f z « 1.0 except z = 0.71 at 3 G for 0.2% sample which is too fast

saturated. The interaction parameter, X = 0.37 for 1 Gauss o f magnetic field strength

can be calculated from Equation 4.4 . The advertised magnetic susceptibility o f 0.8 pm

Bangs 6 8 particles is 1.66x1 O' 2 emu/Oe cm3, which is provided by the manufacturer.

The other interaction parameters for different magnetic strengths are in the Table 4 . 1

(see Appendix B).

94

Figure 4.3 M icroscopy observation for the evolution o f the magnetic latex particle (0.8 pm ) under the magnetic field, 0.2%, 6 G, (a) 0 time, (b) after 5 s.

(fig. con'd.)

95

20 jJtm

(c) after 30 s, (d) after 50 s.

Clus

ter S

ize/

jam

96

5 0

40 -

O

A

i 1 r

1.5A (9 G)

0 .5A ( 6 G)

0 .2A (3 G)

30 -

2 0 - § §

10 - na o

o 4

o

n

A

2 0

A A A A A

j _____ i_

40 60 80 100

T im e /s

Figure 4.4 Cluster size growth for 0.2 wt% o f MLP under different magnetic fields.

Clus

ter

Size

/jnm

97

15 T□ 1.5A (9 G)

O 0.5 A (6 G)

A 0 .2 A (3 G)

10

n n n

n o a

XJm

o

A

0 50 1 0 0

T im e /s

150

A

200

Figure 4.5 Cluster size growth for 0.1 wt% o f MLP under different magnetic fields.

in (s

(t)-

l) In

(s(t

)-l)

98

9 G 0.98+/-0.09

6 G 0.93+/-0.11

3 G 1.0+/-0.203

2

1

0.71+/-0.03

0

0 1 2 3 4

In t

9 G 1.09+/-0.08

6 G 1.06+/-0.12

2 3 G I.06+/-O.I5

1

0

1

Figure 4.6 Pow er law growth for (a) 0.2 w t% (b) 0.1 w t% o f M LP under different magnetic fields, x is a constant for each different field.

99

Table 4.1 The relation between the magnetic strength (Gauss) and scaling parameters. Interaction parameter: X, sample response to the magnetic field: 1/x, and the scaling constant: z.

Concentration Mag. Field

StrengthX 1 / t

( 1 /s)

z

3G 3.33 0.13 1 .0 0 ± 0 . 2 0

(0.71±0.03)

0 .2 % 6 G 13.3 0.64 0.93±0.11

9G 30.0 1.09 0.98±0.09

3G 3.33 0.044 1.06±0.15

0 . 1 % 6 G 13.3 0.059 1.08+0.12

9G 30.0 0.103 1.09+0.08

Our result corresponds approximately to the y = 0 limit by the theoretical

approach o f Miyazima et al. For the case y = 0 (z = 1), which means mass independent

diffusion coefficient, time increased uniformly as 1/N(t) where N (t) is the total number

o f clusters at time t. In case o f y ^ 0, clusters are represented by a random walk on a

lattice. Under no magnetic field condition, the particles are randomly distributed and

represent diffusion kinetics. Under a homogeneous magnetic field whose field energy is

much bigger than kT, the particles come together and the cluster size is increased

uniformly. The strong interaction in the direction o f orientation makes it possible to

neglect the diffusion process in M LP or ER fluid.. Our study shows a faster time scale

than that o f Fermigier et al. and Fraden et al. [4.4 and 4.11] One possible explanation

for the faster aggregation process in our experiment is the selection o f the initial

aggregation time region, which is chosen well below the saturation point o f the

clusters. The clusters growing process slows after saturation begins, and diffusive

motion o f the clusters appears. Coarsening processes that normally occur in ER- or

100

ferro-fluid under the high field strength couldn't be observed with our experimental

conditions, A. < 30, and <{>e ~ 0.06.

4.4 VIDEO SMALL ANGLE LIGHT SCATTERING

4.4.1 GENERAL INTRODUCTION

Video small angle light scattering (VSALS) uses linearly polarized light to

illuminate the object under examination to produce a scattered intensity pattern on a

screen. The scattered pattern is normally at a small angle from the incident light, and

the image is recorded by the video camera and saved on video tape; it was recorded on

photographic plates before the ascendancy o f video technology [4.32], VSALS

patterns from specimens have been used to study the morphology o f crystalline

polymers in the solid state [4.33], which gives the size, shape and orientation o f

molecular axes in such crystalline micro structures [4.34] Typical SALS patterns, for

example clover shape lobes, have been explained by orientation fluctuations in

crystalline and semi-crystalline polymers. Stein et a l [4.33] provided the theoretical

approach based on the classical light scattering approximation and it was proved

experimentally.

Orientation fluctuations, arising from heterogeneity o f the local degree o f

crystallinity, are the origin o f the small angle scattering pattern. Several SALS

experiments have been attempted to follow the melting and crystallization o f the

polymer [4.35], VSALS had been used to investigate the electrorheological fluids in

1992 [4.36], After the application o f the electrical field to the ER fluid, the aligned

clusters altered the properties o f the solution. The aligned clusters caused the solution

to behave more like a solid. The small angle scattering pattern arose from orientation

and surface structure o f the ER fluid. Refractive index matching between the particle

and solvent is an important factor during a VSALS experiment, especially for the high

loading o f solids or for thick samples. However, multiple scattering was not a serious

101

problem in our experiments which are for dilute particles in these cells. Stray light was

reduced by adding w ater between the spacing o f the Vitro cell and the Helmholtz coil.

Halsey [4.37] explained the structure formation in E R fluids as having two

distinct phases on two different time scales. An aggregation time scale, ta, is the

duration to make the initial chains or clusters. Another time scale, a coarsening time,

tc, is the time required to draft each chain and leads to a coarsening o f the fibrous

structure. Normally The typical aggregation time scale is due to a balancing o f

the applied force against the viscosity o f the solvent, ta ~ rj J(E o r M )2; here r | 0 is

solvent viscosity and E or M is the applied electric or magnetic field, respectively. The

kinetic study o f VSALS patterns provides a statistical evalution o f the geometry o f the

scattering entities which is difficult to achieve using optical microscopy. VSALS in this

research is focused on the structure formation time scale, ta.

4.4.2 EXPERIMENTAL SETUP

The VSALS apparatus and its scale are shown in Fig 4.7. A red (670 nm) or

green (543 n m ) laser beam is passed directly through a polarizer and the sample which

is under the magnetic field with Helmholtz coil geometry (see the cell geometry o f

video optical microscopy experiment, Section 4.3.2 ).

The length-scale regime is from 2k /q = 2.5 to 43 pm, where q = 4 7 in sin(0/2)/^o

is the scattering vector magnitude, 0 is the internal scattering angle, n is the refractive

index, and XQ is the wavelength o f the incident light. The value o f 0 was corrected by

applying Snell's law at the water-glass-air interface o f the scattering cell in which the

water-glass interface is negligible and sin ©external= 133 sin ©internal- W hen we

consider the thickness o f the cell, 1 mm, the deviation o f the length scale is within 1 %.

On the imaging screen the intensity, I ~ 1/L2 where L is the distance from the

center to the certain point on the imaging screen, should reduce to 13% o f the centered

light at the high q and 1 % at the low q.

102

\ Z 7VCR

TV

Mac.Imaging

A nalysis* Excel* Passage* Origin

Imaging Plate

Sam ple/M ag. Field

6.5 cm

0.15 cm

7.5 cm

Power Supply

Lens, Filter, or Polarizer

Laser[670, 543 nm]

Figure 4.7 Video small angle light scattering setup.

103

The corresponding length scale is 0.32 pm and 0.43 pm at high and low q, respectively.

A Kepco bipolar operational PS/Amplifier is used as a current supply. A maximum o f

1.5 A o f electric current is applied to the Helmholtz coil and the induced magnetic field

strength is 24 G. The resulting scattering pattern is shown on white paper in a dark

room and the image is recorded by the Javelin MOS (metal oxide semiconductor) solid

state camera fitted with a Nikon optical lens (AF Micro, 1: 2.8).

Every device is adjusted on the optical rail and one can change the position o f the

device at any time. The image processing hardware and software for the VSALS

experiment is the same as that used with the optical microscope.

4.4.3 DATA H A N D LIN G AND E X PE R IM E N T A L R ESU LTS

The same concentrations (0.2%, 0.1%, and 0.05%) as used for the microscope

have been used for the SALS experiment. Scattered intensity is varied from 0 (bright)

to 255 (dark) gray scale and the slice intensities o f each lobe are measured. Sixteen

(16) consecutive pictures are selected from each sample and Figure 4.8 shows 6

consecutive pictures from the video printer. Figure 4.9 is an example o f the intensity

profile o f the 0.2 wt% sample under the 24 Gauss magnetic field. High MLP

concentrations shows a clear lobe shaped pattern at the beginning. As time proceeds,

the lobe becomes smaller, sharper, and the maximum intensity point moves to low q

(high length scale). The intensity profile o f the low concentration samples, however,

doesn't show the clear lobe shape (Figure 4.10). To apply the scaling concept to the

scattered intensity change, the average o f the left and right side o f the intensity profile

is chosen and plotted on the dimensionless axis (q/qmax, I/Imax)> Figure 4.11. The data

from the left shoulder is fitted by slopes o f 3 + 0.3, and the q ' 3 scaling o f the high q

region shows the two dimensional Porod's law which indicates a sharp surface parallel

to the field. The amplitude o f scattered light is proportional to the Fourier transform o f

the refractive index profile.

104

Figure 4.8 VSALS observation for the evolution o f the magnetic latex particle (0.8 pm) under the magnetic field, 0.2%, 24 G, (a) background intensity (b) after 0.5 s (c) after 1 s

(fig. con'd.)

(d) after 4 s, (e) after 1 0 s (f) after 2 0 s.

Inte

nsity

In

tens

ity

Inte

nsity

In

tens

ity

106

250

200

150

100

0 sec

200 300 400 500 6000 100

250

200

150

100

2 0 sec

600300 400 5000 100 200

250

200

150

100

5040 sec

500 6000 200 300 400100

250

200

150

100

1 1 0 sec500 6000 300 400200100

Arbitrary Number

Figure 4.9 Raw intensity profile o f SALS o f MLP under the magnetic field, 0.5 A/ 6 G, 0.05 wt% , 670 nm.

Inte

nsity

In

tens

ity

Inte

nsity

In

tensit

y In

tens

ity

107

130

0 sec

100 200 300 400 6000 300

230

200

130

100

0.5 sec30

200 400100 300 300 6000

100

1.0 sec

200 300 400 300 6000 100

200

130

100

30

1.5 sec0

600200 300 400 3000 100

250

3.0 sec

600200 300 400 3000 100

Arbitrary Number

Figure 4.10 Raw intensity profile o f SALS o f MLP under the magnetic field, 0.5 A / 6

G, 0.2 wt% , 670 nm.

toO

/Im

ax

>

108

0.0-

<6

slope = -3.0 +/- 0.3

8.5 sec6 . 0 sec3.5 sec1.5 sec1 . 0 sec

ln W w

Figure 4.11 ln(I/Imax) vs. ln(q/qmax) plot. Background intensity subtracted, length scale = 27t/q.

109

I f the refractive index (electron density for X-ray) profile is sharp at the interface, the

intensity is given by Porod's law. In three dimensional Fourier transforms, Porod's law

is I(q) = Kp/q4 where I(q) is scattering intensity, Kp is the Porod constant, and q is the

scattering vector. The generalized Porod equation is described by I(q) = Kpq-(d+1)

where d is the dimension o f the system.

By changing the magnetic field and concentration o f MLP, similar types o f

intensity profiles are obtained. The cluster size is calculated from the following

formula, s(t)=2^/qmax(t). At low MLP concentration it was hard to distinguish the

fimax P°*nt on the intensity profile. The rate o f the cluster size growth is calculated

from this SALS experiment and it is compared with microscopic observation in Table

4.2

Table 4.2. Comparison o f the rate o f cluster size growth between SALS and microscopic observation. Unit: pm/s. It is hard to get the data from the conditions where the time scale is too fast (a) or too slow (b).

X SALS M icroscopy

0.2wt% MLP

8 G 23.7 0 .6 6 +0 . 1 0 0.54±0.05

16 G 94.7 1.60+0.30 1.4010.14

24 G 213.1 1.83+0.30 (a)

0.1 wt% MLP

8 G 23.7 0.21+0.04 0.15+0.02

16 G 94.7 0.30+0.06 0.3210.03

24 G 213.1 0.61+0.10 (a)

0.05wt% MLP

8 G

(b)

(b)

16 G 0.1310.01

24 G (a)

The magnetic field strengths in the middle o f the Helmholtz coil were 8 G, 16 G, and

24 G. Each rate, s'(t), was calculated from the initial slope, 4-8 points, o f the length vs.

time plot. It was too fast to separate the time scale at high magnetic strength, > 20 G.

At low concentrations it was hard to analyze the particle size because o f the small

110

number o f particles and the large error bars. There is good agreement for the rate o f

cluster growth between these two methods.

4.5 SUMMARY

The applied field and concentration dependence o f the magnetic latex particles'

kinetic growth was traced by microscopy and small angle light scattering methods. The

high magnetic susceptibility o f the MLP explains why they align at very weak magnetic

fields. Scaling concepts were applied for this cluster size growth mechanism and their

rate o f cluster formation was compared with SALS and microscopy techniques. Good

agreement was obtained between the two methods, and s(t) ~ t 1 0 was observed during

the initial growth o f the cluster.

The aggregation kinetic is not fully discussed in this Chapter. However, the fast

cluster formation is experimentally investigated under the low magnetic field strength,

low X, where the aggregation is not disrupted by the saturation or the coarsening o f the

structure.

Ill

4.6 REFERENCES

4.1 P. Meakin, Physica Scripta 64, 295 (1992).

4.2 ER Fluid

(1) General

(a) R. Tao, "Electrorheological Fluids," Proc. Inter. Conf. on ER fluids, World Scientific, NJ, 1991.

(b) A. P. Gast and C. F. Zukoski, Advances in Colloid and Interface Sci. 30, 153, (1989)

(c) T. C. Halsey and J. E. Martin, Scientific American 58, Oct. (1993).

(d) T. C. Halsey, Science 258, 761, Oct.(1992).

(e) P. M. Adriani and A. P. Gast, Faraday Discuss. Chem. Soc. 90, 17 (1990).

(2) Properties

(a) H. Uejima, Japn. J. Appl. Phys. 11 (3), 319(1972).

(b) L. Marshall, C. F. Zukoski, J. W. Goodwin, J. Chem. Soc. Faraday Trans. I 85(9), 2785 (1989).

(c) W. S. Yen and P.J. Achom, J. Rheol. 35(7), 1375 (1991).

(d) J. C. Hill and T. H. van Steenkiste, J. Appl. Phys. 70(3), 1207 (1991).

(3) Applications

(a) D. L. Hartsock, R. F. Novak, and G. J. Chaundy, J. Rheol. 35(7), 1305 (1991).

(b) D. A. Brook, Devices Using ER Fluids, Proc. 2nd Conf. on ER Fluid, Raleigh, NC, Aug., 1989.

4.3 Conference on Electrorheological Fluids, Carbondale, EL, R. Tao, Ed., World Scientific, Singapore, 1992

4.4 S. Fraden, A. J. Hurd, and R. B. Meyer, Phys. Rev. Lett. 63(21), 2373 (1989).

4.5 (a) M. Ozaki, H. Suzuki, K. Takahashi, and E. Matijevic, J. Coll. Inter. Sci.113(1), 76(1986).

(b) P. C. Scholten and D. L. A. Tjaden, J. Coll. Inter. Sci. 73(1), 255 (1980).

(c) A. T. Skjeltorp, J. Appl. Phys. 57(1), 3285 (1985)

4.6 P. G. de Gennes and P. Pincus, Phys. Kondens. Mater. 11, 189 (1970).

112

4.7 G. Bossis, C. Mathis, Z. Mimouni, and C. Paparoditis, Europhys. Lett. 11, 133(1990).

4.8 P. C. Jordan, Molecular Phys. 25(4), 961 (1973).

4.9 D. A. Krueger, J. Coll. Inter. Sci. 70(3), 558 (1979).

4.10 C. F. Hayes and S. R. Hwang, J. Coll. Inter. Sci. 60, 443 (1977).

4.11 M. Fermigier and A. P. Gast, J. Coll. Inter. Sci. 154(2), 522 (1992).

4.12 H. Bogardus, D. A. Krueger, D. Thompson, J. Appl. Phys. 49, 3422 (1978).

4.13 K. Sano and M. Doi, J. Phys. Soc. Jpn. 52, 2810 (1983).

4.14 B. B. Mandelbrot, "The Fractal Geometry o f Nature, Freeman," San Francisco, 1982.

4.15 D. A. Weitz, M. Y. Lin, and C. J. Sandroff, Surface Science 158, 147 (1985).

4.16 D. W. Schaefer, Science 243, A1023 (1989).

4.17 P. Meakin, Phys. Rev. Lett. 51, 1119, (1983).

4.18 M. Kolb, R. Botet, and R. Jullien, Phys. Rev. Lett. 51, 1123 (1983).

4.19 S. Miyazima, P. Meakin, and F. Family, Physical Review A 36(3), 1421 (1987).

4.20 J. J. M. Janssen, J. J. M. Baltussen, A. P. van Gelder, and J. A. A. J. Perenboom, J. Phys. D: Appl. Phys. 23, 1447 (1990).

4.21 A. Watanabe, Y. Yamaoka, E. Akaho, K. Kuroda, T. Yokoyama, T. Umeda, Chem. Pham. Bull. 33(4), 1599608 (1985).

4.22 F. V. Geel, B. W. Smith, B. Nicolaissen, J. D. Winefordner, J. Microsc. 133(2), 141 (1984).

4.23 J. A. Manson, A. Ramirez, R. W. Hertzberg, Polym. Mater. Sci. Eng. 50, 106 (1984).

4.24 H. Ambar, Y. Aoki, N. Takai, and T. Asakura, Appl. Phys. B. 38, 71 (1985).

4.25 P. Prentice, S. Hashemi, J. Mat. Sci. 19(2), 518 (1984).

4.26 S. N. Jabr, Opt. Lett. 10(11), 526 (1985).

4.27 S. Inoue, J. Cell Biol. 89, 346 (1981).

4.28 D. L. Commare, American Lab. 20(12), 56 (1988).

4.29 J. D. Cutnell and K. W. Johnson, "Physics," John Wiley & Sons, 1989.

113

4.30 M. Fermigier and A. P. Gast, J. Coll. Inter. Sci. 154(2), 522 (1992).

4.31 S. Fraden, A. J. Hurd, R. B. Meyer, Phys. Rev. Lett. 63, 2373 (1989).

4.32 R. S. Stein and T. Hotta, J. Appl. Phys. 35, 7 (1964).

4.33 (a) R. S. Stein and M. B. Rhodes, J Applied Phys. 31(11), 1873 (1960).

(b) T. Takebe, T. Hashimoto, B. Ernst, P. Navard and R. S. Stein, J Chem. Phys. 92(2), 1386 (1990).

4.34 J. M. Haudin, "Optical Studies o f Polymers," Chapter 4, G. H. M eeten Ed., Elsevier, N. Y., 1986.

4.35 T. Kyu, K. Fujita, M. H. Cho, T. Kikutani, and J-S. Lin, Macromolecules 22, 2238 (1989).

4.36 J. E. Martin and J. Odinek, Phys. Rev. Lett. 69(10), 1524 (1992).

4.37 T. C. Halsey and W. Toor, Phys. Rev. Lett. 65(22), 2820 (1990).

CHAPTER 5

DYNAMIC BEHAVIOR OF A HIGH STRENGTH ROD-LIKE POLYELECTROLYTE IN A STRONG ACID

114

115

5.1 INTRODUCTION

The behavior o f polyelectrolyte solutions, which consist o f polyions (charged

polymer chains), counter ions, and coions, such as Na+, Cl", H+, is one o f the most

controversial issues o f polymer physical chemistry. Theoretical treatm ent based on the

Poisson-Boltzman equation o f electric potential leads to predicted phenomena like

small-ion polyion coupling modes [5.1], counterion condensation [5.2], electrolyte

dissipation [5.3], and scaling laws [5.4], Many different experimental techniques have

been applied to polyelectrolyte systems: static and dynamic light scattering [5.5], small

angle X-ray or neutron scattering [5.6], electrophoretic light scattering [5.7],

viscometry [5.8], NM R [5.9],

The dynamics o f polyelectrolytes are determined mostly by electrostatic forces

between the charges along the polymer chains and low molecular weight ions. At the

high salt condition, polyelectrolytes behave like normal, uncharged polymers. Unusual

and unexpected properties o f polyelectrolyte solutions have been reported at the low

salt condition. One o f them is "ordered structures." The notion o f ordered structures

is based on the SAXS data o f tRNA, BSA [5.10], and N aPSS[l 1] which have a single

diffuse "Bragg-like" peak. This peak disappears at the isoelectric point, which implies

the ordered structure is caused by electrostatic interactions. Another example is the

extraordinary-ordinary phase transition, "O-E or E -0 transition"', at a critical salt

concentration, the diffusion coefficient dropped catastrophically by over an order o f

magnitude. Lin et al. first reported the O-E transition o f poly(L-lysine)/KBr system

[5.12], Since then, many different polyelectrolytes have been found that exhibit the O-

E transition, including NaPSS [5.13], DNA [5.14], bovine serum albumin (BSA)

[5.15]. In the present study possible ordered structure and its relation to the O-E

transition have been investigated for the first time by dynamic light scattering with rigid

backbone polyelectrolytes. This type o f polymer emphasizes intermolecular interaction

116

rather than intramolecular interaction since the chains are thought to be almost fully

extended even at high salt content [5.16],

Synthetic rod-like polymers that have high strength and thermal stability are

increasingly being applied in polymer industries such as aerospace and electronics.

Poly(p-phenylene cis benzobisoxazole) (PBO: Figure 5.1) is one o f the rigid rod

polymers having a specific tensile strength 10 times that o f steel [5.17], It is only

soluble in strong acids such as sulfuric acid, methane sulfonic acid (MSA) and

chlorosulfonic acid. These acids protonate the nitrogen or oxygen sites o f the polymer

backbone and the polymers become polyelectrolytes [5.18], Thus, it is very important

to consider the presence o f salt and the ionic strength o f the solvent when we

investigate the solution properties o f PBO.

Berry and co-workers showed the rod-like nature [5.19] and electrostatic

effects [5.16] o f high strength polymers such as PBO and PPTA (poly 1,4 p-phenylene

terephthalamide). At low ionic strength, the polymer solutions had high viscosity ,

small translational diffusion coefficients, and decreased unpolarized scattering. They

proposed intermolecular ordering due to the electrostatic repulsion between protonated

polymer chains. Berry et al. believed these interactions had little effect on chain

conformation, owing to the large persistence length obtained at moderately high ionic

strength. Schmitz [5.20] discussed the possibility o f a O-E transition o f Berry's PBO,

comparable to that o f poly(L-lysine). H e argued that there was "temporal aggregation"

in the extraordinary regime. Berry et al. also have suggested the formation o f

paralleled dimers, trimers, etc. (aggregation) as intermolecular repulsions become

screened out by counterions when salt is added.

Fifteen years after the Berry studies, Roitman, working at Dow's W alnut Creek

facility, did static light scattering and viscosity measurements on PBO.

117

n

1.21nm

Figure 5.1 Chemical structure o f cis-PBO. At least tw o sites o f N or O are protonated by strong acids. Length scale is from K. Tashiro and M. Kobayashi, Macromolecules 2 4 ,3706(1991).

118

4.0

C(PBO) = 49.7 g/ml 2 g . Solvent: r| = 14.56 cP

o

o2.5-| O

o oo

2.0o o o

0.00 0.05 0.10 0.15NaCH3S 0 3 Conc.(M)

Figure 5.2 Intrinsic viscosity o f PBO/M SA-M SAA as a function o f salt. Data file is provided by Dr. D. B. Roitman at Dow's Walnut Creek group in California.

119

This differs from the Berry study in that PBO [5.21], not PBT was studied. Also acid

anhydride was added to reduce moisture. The high intrinsic viscosity in the absence o f

salt drops sharply with added salt and reaches a "plateau" at higher ionic strength

(NaMSA>0.01 M : Figure 5.2). The scattering from PBO solutions at low ionic

strength is rather anomalous: low intensities (despite dn/dc = 0.47 ml/g) and very

rapidly decaying correlation functions are observed. The viscosity behavior o f PBO is

much more sensitive to salt than is that o f poly(L-lysine). This phenomena was

explained by tw o possibilities. The first possibility is that intramolecular interactions

among charged sites on the chain may become screened by addition o f the salt.

W ithout such interaction, the chain may assume a less extended conformation. In

contrast, Berry felt that the chains were very stiff already and could not be further

extended. He explains the reduction in [r|] through intermolecular associations- i.e.

side-by side aggregation o f rods into bundles of, perhaps, 2-3.

Dynamic light scattering (DLS) is a very challenging experiment for the

PBO/M SA-M SAA polyelectrolyte system. Preliminary dynamic light scattering

experiments by M. Tracy [5.22] show correlation functions measured from PBO

solutions at low ionic strength, 0.01 M NaM SA, have very fast decays and low

intensities. Tracy did not find a very well behaved dynamic signal from the very fast

decay. At high ionic strengths, 0.05 M NaMSA, the scattering appears "normal."

In addition to the polarized scattering o f the rigid rod-like polymer, depolarized

scattering originating from the optical anisotropy o f the rod is possible -- but the

experiments are again challenging [5.23]. The decay o f the correlation function o f the

depolarized scattering is dominated by the rate from rotation which is ordinarily much

faster than the rate from translation ( 6 Dr > q2 D J. Fabry-Perot interferometry can be

used for the fast rotational diffusion studies [5.24], Berry et al. reported that the

depolarized intensity decreases with decreasing salt concentration in solutions

120

containing the PBT type rod. Depolarized scattering o f PBT also has been measured

by Crosby et al. Their point o f view was to characterize its conformation and to

estimate the polydispersity o f a polymer sample. Extensive measurements, in the

present work, have been conducted using the ALV-5000 correlator with several

different salt and concentration ranges. The ALV-5000 is a vastly superior correlator

to those o f the Berry [5.21] or Crosby [5.23] studies. Special emphasis is devoted to

the decay mode at low salt condition. Even though no definite interpretation is yet

available, there are several interesting observations, such as the absence o f the ordinary

O-E transition and the presence o f an undefined fast decay mode at low salt condition,

in both polarized and depolarized scattering.

5 .2 E X P E R IM E N T A L

5.2.1 M A TE R IA LS

The PBO were provided by Dow's Walnut Creek group in California, and each

solution was prepared by Dr. D. B. Roitman. The solvent and solutions were subjected

to N 2 bubbling before filtering, and were filtered under an N 2 blanket. A 0.2 pm PTFE

filter was used for filtering and the sample was saved in 1 0 mm clean test tubes, which

had been dedusted and tested at LSU and then sent to Dow for loading. Every sample

cell was sealed by a PTFE - faced screw cap, wrapped with aluminum foil and kept in a

desiccator with N 2 atmosphere before and after measurement to prevent oxidation and

photo-degradation. Table 5.1 summarizes the samples, which differ by molecular

weight, concentration, and salt condition o f PBO/MSA-MSAA. Commercial M SA

was distilled and standardized by saturating the M SA fractions with M SAA to control

the w ater content. The strongly corrosive nature o f solvent requires special handling

precautions and details appear elsewhere [5.25], Length scales o f the polyelectrolytes

also appear in the table. Relatively high molecular weight PBO samples (approximate

numbers M = 78,000, [r|] = 2,000 ml/g, L = 400 nm & L/Q = 13, where L is the

121

contour length and Q is the persistence length) and low molecular weight samples (M

= 19,000, [rj] = 540 ml/g, L = 98 nm & L/Q = 3-4), and sets o f polymer concentrations

(0, lx lO -4, 5x10^, 1.5xlO'3, 2 .5 x l0 '3, 4 x l0 " 3 g/ml), and also complete sets o f salt

concentrations (0, 0.01, 0.1, 0.45 M NaM SA) were prepared for the DLS studies.

Some samples discolor and acquire a bluish tone or have white precipitates due

to water contamination during storage. The discoloration did not affect the molecular

weight, just diminished the intensity o f the scattered light. The cloudy and whitish

samples were avoided. Even after centrifuging the samples, the dust level at low

concentration and low salt condition was still too high to obtain a quiet DLS signal.

Therefore, the experiments focused on the high concentration samples.

Table 5.1 Samples o f PBO/M SA-M SAA with different concentrations and salt conditions. Intrinsic viscosity was measured with solvent viscosity, 14.56 cP .without salt (NaMSA), at 25°C.

M olecular Weight 19,000 78,000

Intrinsic Viscosity 540 ml/g 2 , 0 0 0 ml/g

Contour Length 98 nm 400 nm

Concentration lxlO "4 g/ml 5x10"4 g/ml lxlO " 3 g/ml 4x10 ' 3 g/ml

Salt Condition 0 M 0.01 M 0.10 M 0.45 M

5.2.2 M EA SU R E M E N T

Some preliminary measurements were performed using X0 = 632.8 nm He-Ne

laser to circumvent the fluorescence problems associated with shorter wavelengths.

W e obtained essentially the same results with a pow er controllable ion laser at either X0

= 488.0 nm or 514.5 nm. To eliminate the fluorescence intensity when we used the X0

= 488.0 nm or 514.5 nm laser, we inserted a narrow band optical filter in front o f the

photo detector. The laser beam passed through the vertical polarizer, a low-power

122

Glan-Thomson polarizer (Karl Lambrecht MGT3E5; 1W cm"2), and was narrowed by

using a lOx microscope objective. The focused beam decreases the diameter o f the

scattering region which helps avoid dust and also increases the size o f the coherence

area. High coherence area allows one to maximize the power per coherence area. This

narrow beam illuminates the sample which was held in a refractive index matching bath

o f toluene. A neutral density filter is used to greatly reduce the back-reflected beam at

the air-glass interface o f the matching bath. The scattered light passed through an

analyzer and then through the optical filter having the same wavelength as the incident

light. Polarized and depolarized measurements were done by changing the angle o f the

polarizer and analyzer (details in Chapter 2). These signals were collected by a EM I-

9863a PM tube, passed through the amplifier/discriminator, and sent to the ALV-5000

correlator. Samples which had some dust were centrifuged several hours at 5,000 rpm

(4800 g, g is proportional to co2 r, where co is the angular speed and r is the radius, « 10

cm ). Several hundred runs were collected to increase the acquisition time and were

selectively averaged with new software called ALVAN. ALVAN is an in-house

developed program for correlation function analysis for data gathered from the ALV.

It enables cumulant fitting, multiple summing, non-ergodic analysis, etc. Figure 5.3

compares a single run and an average o f 50 runs. A single 30 s measurement exhibits

high frequency noise, but in the sum o f 50 runs, the high frequency portion has become

quiet enough to obtain accurate results.

5.3 RESULTS

5.3.1 SIG N A L C O N SID ER A TIO N

At high salt and high polymer concentration, the scattering signal is relatively

strong and the autocorrelation function is easily measured.

123

1.08

1.06

1.04

1.02

1.00

1 0 "* 1 0 1 1 0 ° 1 0 1

t / s

1.08

1.06

1.04CN

1.02

1.00

1 0 11 0 °

x /s

Figure 5.3 D ata handling with single short run and multiple runs. Single run has high frequency noise at short lag time, but multiple runs have quiet enough to measure a very rapid decay. Data was analyzed by the in house developed program ALVAN.

124

Polarized (Vv) correlation functions, which give the same result as unpolarized

correlation function (Uv), for the high salt solution (M = 19,000, [r|] = 540 ml/g,

4 x l0 ' 3 g/ml conc., and 0.1 M salt and 0.45 M salt) have been examined (Figure 5.4(b)).

Two decay modes are seen, a dominant slow decay mode in the range o f 1,000-5,000/s,

and fast mode, « 100,000/s, as shown in Figure 5.4(a). The correlation function

measured in this configuration shows decreased coherence (i.e. the coherence factor

f(A) is ~ 0.25). The autocorrelation function o f normal latex particles, diameter =

0.087 pm, was measured to determine the maximum f(A) at the same geometry, same

pin-hole setting (front 800 pm and rear 200 pm) and same wavelength (488 nm) in

Figure 5.4(c). The value o f f(A)max was about 0.32. In both the latex and high salt

PBO/M SA-M SAA cases, the scattering by solvent is negligible. The slightly reduced

f(A) in the PBO/M SA-M SAA probably represents slight beam expansion due to the

thermal lensing effects. It is not possible to rigorously exclude the possibility o f very

fast decay modes lying outside the correlator window, but the short time data o f Figure

5.4(a) do appear to have almost reached a plateau. Figure 5.4(c and d) shows clear

single exponential decay and reaches f(A) ~ 0.32 which is around 20% larger than that

o f the PBO sample.

Figure 5.5(a) shows a sample correlation function at the low salt condition,

0.01 M , M = 19,000, 1,000 s acquisition time at 90° scattering angle, and same pin­

hole setting as previously. The scattered intensity o f the solution was very weak, only

around 2 times bigger than that o f solvent, and the correlation function has low f(A),

low coherence. The signal was verified by measuring the weak scatterer, low

concentration PS/toluene. Figure 5.5(b) shows the autocorrelation function o f

polystyrene 5x1 O' 4 g/ml, M = 422,000, in toluene at 25°C, 90° angle, and the same pin­

hole setting (800/200) as PBO/M SA-M SAA measurement.

125

m ax

1. 0 -

TTTT

t / s0.0

-0 .1 -

-0.2 -

- 0 .3 -

-0 .40 .0 0 0 .0 5 0 .1 0 0 .1 5 0 .2 0

t/ 1 0 ' 3s

Figure 5.4(a) Correlation function for PBO/M SA-M SAA solution: M = 19,000; Cone. = 4x10 ' 3 g/ml; cs = 0.4 M; Vv measurement; 90° scattering angle; 488 nm wavelength; 800/200 pin-hole setting; 300 s acquisition time single run; 25°C. (b) corresponding semilog representation.

126

1 .2 -

cs'"on

1 .0 -

1 0 '

0 . 0

-i-

- 0 .2 -

-2 -

0.50 . 0

c -0.4-

- 0 . 6 -

0 . 0 0.40 . 2

t/ 1 0 ’3s

Figure 5.4(c) Correlation function with strongly scattered latex, 0.087 p.m, solution under the same pin-hole setting (800/200), X = 488 nm. (d) corresponding semilog representation.

127

csCd)

1.3-

1 .2 -

1 . 1 -

1 .0 -

-—

° □°u°Drf,o,

X.(a)

i i 11 m i f - T T 11 l in y

1 0,-7

1 0 H 1 0 -| i v rTivii| i i i ivvii | v i i i

1 0 " 1 0i-3

1 0r2

1 0-1

1 0 °

x/s1 0 1

CM

bD

f(A)1.3 max

1 . 2

1 . 1

1 . 0

1 0 °

x/s

Figure 5.5(a) Correlation function for low salt condition o f PBO/M SA-M SAA solution: M = 19,000; Cone. = 4x1 O' 3 g/ml; cs = 0 M; Vv measurement; 90° scattering angle; 488 nm wavelength; 800/200 pin-hole settings; 1,000 s acquisition time single run; 25°C. (b) The comparison o f correlation function with low f(A) polystyrene (M = 422,000) in toluene under the same pin-hole setting (800/200), 90° scattering angle; 488 nm wavelength.

128

The intensity ratio between PS and toluene is adjusted to be close to the same ratio as

PBO and MSA, Ips/toi^tol= IpBO^soivent ~ 2 which gives the similar f(A) value as the

low salt PBO sample. PS/toluene with low f(A), 0.2, gives a clear correlation function

plateau even despite the low coherence. As the low time value o f g(2M approaches

f(A)max> there is probably not too much rapidly decaying signal that has been missed in

the PBO/M SA low salt samples. However, future studies would benefit from a

correlator capable o f measuring at shorter decay times.

5.3.2 SIMPLE ANALYSIS

5.3.2.1 POLARIZED MEASUREMENTS

A series o f autocorrelation functions for high molecular weight (M = 78,000,

[r|] = 2,000 ml/g in 0.1 M NaM SA, 4 x l0 " 3 g/ml conc.) is shown in Figure 5.6(a). The

bimodal, or possibly more complex, nature o f the decay at high salt is clearly evident

from the semilogarithmic representation in Figure 5.6(b). As salt concentration

decreases, the dominant slower mode disappears while the fast mode remains in the

polarized scattering. The correlation function for the solvent itself was measured with

3,000 s sample time (400 runs @ 20 s, keeping « 40% o f the runs) and is shown in the

same plot. It is clear that the weak fast mode cannot be attributed to the complex

nature o f the solvent, or to afterpulsing or other common short-time artifacts. When

the concentration decreased from 4x10 ' 3 g/ml to IxlO ' 3 g/ml (not shown), trends o f the

slow and fast mode decay, slow and fast mode at high salt and only fast mode at low

salt condition, are the same.

Figure 5.7 is the same type o f plot as Figure 5.6 but for the low molecular

weight samples. The top plot has already appeared as Figure 5.4 (a). A slight

improvement in coherence over that in Figure 5.6 is due to the increased power o f the

laser and smaller pin-hole settings.

129

□□ Di1 .2 -

'"'00

x/s

0 0

° 0 .1 0 M1.2

1.1

1.0

x/s

1 .2 -

/w'1 .1 -<N

0 0

1 .0 -

0 0

1 .2 -

1 .1 -

1 .0 -

° 0.01 M

Solvent

1 0 -" 1 0

Figure 5.6(a) Correlation functions for high molecular weight PBO/M SA-M SAA solution and their salt dependence: M = 78,000; Cone. = 4 x l0 " 3 g/ml; Vv measurement; 90° scattering angle; 632.8 nm wavelength; 1,000/100 pin-hole settings; « 1500 s acquisition time with 99 runs; 25°C. Correlation function for the solvent is also shown in the bottom plot.

ln(g

(t))

ln(g

u;(x

))

130

0.0-

+ + + +

- 0 . 1 -* +

0.0

i r ' -o. 2v-6o£

-0.4

++ + +

0.0 0.5 1.0 1.5

t /1 0 ‘3S

- 0 .2 - (b)o .o

T0.1 0.2 0.3

t/ 1 0 ' 3 s

0 . 0

-0.5-

t

- 1.0

-1.5

+ ++ *

+ + +0.0 0.5 1.0 1.5 2.0

t / 1 0 ' 3s

(C )

0 .0 " o T 0.2

x/10"3s

0.3 0.4

Figure 5.6(b) Corresponding semilog plot o f Figure 5.6(a) for 0.45 M salt solution, and (c) for 0 M salt solution.

131

° 0.45M salt

1.0

1CT

1

° O.lOMsaltlH ,

1 0 '5 1 0 '2 1 0 '1

t / si

° O.OlMsaltl

iso l

t/ s

1

° No salt Solvent

Figure 5.7(a) Autocorrelation functions for low molecular weight PBO/M SA-M SAA solution and their salt dependence: M = 19,000; Cone. = 4 x l0 " 3 g/ml; Vv measurement; 90° scattering angle; 488 nm wavelength; 800/200 pin-hole settings; 300 - 1,000 s acquisition time with single run; 25°C. Top panel for 0.45 M salt is same data as Figure 5.4 (a).

132

0.61.0 0.45 M salt 0.50.8 ^ 0.4

S 03 < 0.2

0.1

0.6

< 0.4

0.2

0.0 0.0

r / io V r/io 3 s'.3 „-l

0.61.0 0.10 M salt0.10M salt 0.50.8 f{A) = 0.250.40.6

< 0.40.3

< 0.2

0.10.2

0.0 0.0•3 ■2 ,0 |1 ,4 ■3 •2 .3 ,4,2 |3 ,0

r/ioJ s.3 r/io 3 s'3 „-l

0.61.0

0.01 M salt 0.50.8

7=: 0 40.6

< 0.4S 0.3 x< 0.2

0.2 0.1

0.0 0.0■3 •2 .0 ,0 |3,1 ,2 ,3 |4 |4

r/io 3 s,3 „-l r / io V0.6

1.0No salt 0.5

0.8 0.4/•—s

g 0.3 < 0.2

0.6

< 0.4

0 . 2 0.1

0.0 0.0•3 ■2 ,0 ,4 •3 •2 1 |0 ,2 ,3,2 ,3 |4

r/io 3 s',3 r/io 3 s-1

Figure 5.7(b) Decay rate distributions corresponding to Figure 5.7(a), and amplitude x f(A ) 1 / 2 vs. decay rate plot.

133

By decreasing the salt concentration, the same trends are observed despite the

molecular weight difference (Figure 5.6 is for M = 78,000 and Figure 5.7 is for M =

19,000).

The fitting curve comes from the CONTIN program provided by the ALV-5000

correlator (parameters: fit range = 1 - 2 x l0 7 Hz, # o f bins = 32, Linear Coeffic. = 0,

and Prob. Level = 0.5). Normalized ALV CONTIN data were converted to amplitude

and multiplied by f(A ) 1 / 2 in the same plot. The slow mode disappeared with decreasing

salt and only the fast mode remains in this plot. The converted amplitude o f the fast

mode slightly grows upon decreasing the salt concentration, which means the fast mode

is affected somewhat by dominant slow decay at high salt.

The polymer concentration dependence is not obvious from the CONTIN

analysis, but the measured diffusion coefficient increases with increasing polymer

concentration. The decay rate and corresponding diffusion coefficient as a function o f

concentration are shown in Table 5.2. Each value is from CONTIN analysis o f the

ALV correlator based on the parameters, wavelength: 488 nm, refractive index: 1.43,

viscosity: 22.34 cP for 0.446M salt and 16.29 for 0.10 M salt, temperature: 25°C.

Table 5.2 Slow decay rate and corresponding apparent diffusion coefficient o f the low molecular weight (M = 19,000), at tw o different salt conditions, and 4 different concentrations. Each value is from the CONTIN analysis o f the ALV correlator.0 = 90°, X0 = 488 nm, Decay rate, T : 1/s, Diffusion coefficient Dt: 10' 8 cm 2 /s.

conc.(g/ml)

1 x 1 0 ^ 5x1 O' 4 lx l 0 ’ 3 4x10 ’ 3

0.10 M salt

r - 2 , 0 2 0 2,150 2,280D, - 2.98 3.17 3.36

0.446Msalt

r 998 1,080 1,190 2,400D, 1.47 1.59 1.75 3.53

134

Due to the high dust level in many low concentration samples it was hard to get

the data at low salt and low concentration. A set o f results (which is not in the Table

5.2) shows that the measured diffusion coefficient o f the slower mode o f M = 78,000

PBO/M SA is lower than that o f the M = 19,000 sample, as expected.

5.3.2.2 DEPOLARIZED MEASUREMENTS

W e have seen that polarized measurements give fast and slow mode decays.

The slow mode is the primary mode at the high salt condition and the relative amplitude

o f the fast mode is reduced by adding salt. The fast mode is the dominant mode at low

salt condition.

Figure 5.8(a) shows the autocorrelation function o f the depolarized scattering

o f low molecular weight sample, and Figure 5.8(b) is the corresponding decay rate

distribution. The depolarized measurements o f low and high salt conditions o f

PBO/M SA-M SAA only have a broad fast mode decay, « 100 kHz. It seems that the

decay rate doesn't depend on the salt concentration. The depolarized intensity o f the

solvent, M SA-M SAA, is near the background noise level o f the instrument, and the

depolarized signal ratio o f polymer to solvent is around 2 times bigger than that o f

polarized signal. In fact, the lowest concentration sample has more than 4 times

stronger intensity than does the solvent in depolarized scattering. The polarized

scattering ratio, IVv pbc/^Vv, solvent- is just 1.5 for the same samples.

The initial correlation function at lag time around 10*6 s is still noisy, after the

1 2 0 0 s retained runs.

5.3.2.3 ANGULAR DEPENDENCE

Typical angular dependence o f the autocorrelation function for tw o different

salt and tw o different concentration samples are in the Figure 5.9(a) and (b),

respectively.

° No salt

1.4 oo

--------Solvent

<N"to

°n

1.2 ■

1.0 ... ........... .10'7 10-* 10's 10“* 10° 10'2 10'1 10° 101 10J

t/ s

Figure 5.8(a) Correlation function o f depolarized measurement o f PBO/M SA-M SAA solution and their salt dependence: M = 19,000; Cone. = 4 x l0 " 3 g/ml; H v measurement; 90° scattering angle; 488 nm wavelength; 800/200 pin-hole settings; 1 2 0 0 s acquisition time with single run; 25°C.

1 . 0

0 . 8

0 . 6

0.40 . 2

0 . 0

0.45M salt

>3a l4 ,5 ,6|0 ,1

T / s 1

1 . 0

0 . 8

0 . 6

< 0.40 . 2

0 . 0

0.10M salt

,3 >7|4 ,5,0 |21

r/s'1

1.00.80.6

0.01M salt

0.20.0

fi ,6 J,3 ,4,0 ,2

r/s'1

1.0No salt

0.80.6

^ 0.40.2

0.0

Figure 5.8(b) Corresponding decay rate o f Figure 5.8(a).

137

M easurement was performed at 6 different angles, 30°, 45°, 60°, 90°, 120°, and 135°

at 25°C, = 632.8 nm, using a single run for the high molecular weight (M = 78,000)

samples (Figure 5.9(a) and (c)). For the low molecular weight sample, A,0 = 514.5 nm,

5 different angles, and multiple runs were used (Figure 5.9(b) and (d)). The correlation

function for the scattering angle 90° and 120° in Figure 5.9(d) is not shown because we

changed the pin-hole settings and power o f the laser at this point.

Two data-fitting routines were applied in an attempt to identify decay rates T

and amplitudes A o f those correlation plots. Linear q2 dependence o f the 3rd cumulant

(channel selection = 3 - 128, weight 1) T o f the high molecular weight, M = 78,000,

high concentration (4 x l0 ' 3 g/ml), at various salt conditions, is shown in Figure 5.10.

The behavior in high salt is normal which means there is a zero intercept and finite

slope in the plot o f T vs. q2. The slopes, translational diffusion coefficients, in 0.45 M

and 0.1 M salt solutions are 3 .0 6 x l0 '8 cm2/s and 1.26xl0 '7 cm 2 /s, respectively.

However, the low salt sample shows a huge intercept and deviation from the linear

fitting line. The abnormal behavior at the low salt condition was re-analyzed by two

exponential decay and ALV-CONTIN analysis. When we analyze the data with

CONTIN in the ALV, there are two decay modes. The slow mode decay exhibits

typical diffusive behavior with a positive slope and zero intercept, but the fast mode has

a huge intercept and doesn't depend on the q vector.

A different concentration sample (0.4M NaM SA, 2x10 ' 3 g/ml), which is not

shown in the plot, also gives q-dependence o f the slow mode and q-independence o f

the fast mode. The fast mode in low salt seems q-independent. The angular

independence o f the fast mode is unclear except to note that the fast mode has a huge

intercept.

138

1.30

1 .2 5 -

v jw B1.2 0 -

1201351 .1 5 -

1.1 0 -

1 .0 5 -

1 .0 0 -

t / s

Figure 5.9(a) Angular dependence o f the autocorrelation functions for the high molecular weight PBO/M SA-M SAA solution at high salt condition: M = 78,000; Cone. = 4x10" 3 g/ml; cs = 0.45 M; Vv measurement; 6 different angles; 632.8 nm wavelength; 1,000/100 pin-hole setting; 30 s acquisition time with single run; 25°C, f(A)max = 0.40. (b) low molecular weight sample: M = 19,000; Cone. = 4x10 ' 3 g/ml; cs = 0.45 M; Vv measurement; 6 different angles; 514.5 nm wavelength; 600/100 pin­hole settings; ~ 15,000 s acquisition time with multiple runs; 25°C, f(A)max = 0.70.

139

0 0

1.10

60

120135

1.05

1.0 0 -

t/ s

<N0 0

1 .0 -

x/s

Figure 5.9(c) Angular dependence o f the autocorrelation functions for the high molecular weight PBO/M SA-M SAA solution at low salt condition: M = 78,000; Cone. = 4x10 ‘ 3 g/ml; cs = 0 M; Vv measurement; 6 different angles; 632.8 nm wavelength; 1 ,0 0 0 / 1 0 0 pin-hole setting; 30 s acquisition time with single run; 25°C, f(A)max = 0.40. (d) low molecular weight sample: M = 19,000; Cone. = 4 x l0 ‘ 3 g/ml; cs = 0 M; Vv measurement; 3 different angles; 514.5 nm wavelength; 800/200 pin-hole settings; ~ 15,000 s acquisition time with multiple runs; 25°C, fl^A ),^ = 0.32.

140

10-C/3

of 'o

i- 2

(1.74+/-0.1) x 107

0 - '

-©— ©--A A -(3.47 +/- 0.3) x 10'1

0

-A-

4 6q2/1010cm'2

□ 0.01 M saltO 0.10 M saltA 0.45 M salt

T "8 1 0

□ 0.01 M salto 0 . 1 0 M salta 0.45 M salt

co

int.=6085 +/- 1110 s'1 slope=3.95 +/- 0.25 xlO'7 c m V

1 -. o

int.=558 +/- 220 s'1slope=l .26 +/- 0.05 xlO"7 cm V 1

. , - ' ' e 'int^n+ASO s ' 1

slope=3.06 +/- 0.10 xlO"7 cm2s‘1

q /1 0 cm

Figure 5.10 T vs. q2 plot (bottom) and I7q 2 vs. q2 plot (top) o f the third cumulant o f Vv autocorrelation functions. High molecular weight (78,000), high concentration 4 x l0 ‘ 3 g/ml and different salt conditions (0.01, 0.10, 0.45 M), at 25°C. These data are based on single run with total acquisition time ~ 30 s.

141

The quiet data set o f Figure 5.9(b) and (d) shows the angular dependence o f the

slow mode and angular independence o f the fast mode. This quiet data set is used for

the further analysis.

5.3.3 CONTIN ANALYSIS WITH MULTIPLE RUNS

Previously, we have used CONTIN rather non-critically to provide a sketch o f

the decay profiles. We now take a closer look using CONTIN analysis with multiple

runs and multiple angles. The relation between normalized correlation function, g(x),

and distribution function, G(T), is a severely "ill-posed" problem. This means noise

contributions lead to a typically infinite set o f different distributions G(T) that are

consistent with the measured g(x). CONTIN is a package for inverting noisy linear

operator equations originated by S. Provencher. The basic principles o f ALV CONTIN

and the original CONTIN by Provencher, PRO CONTIN, are the same. However, the

tw o programs use a number o f different quadrature settings and different maximum bin

numbers (ALV's = 44, Povencher's = 6 6 as implemented in this laboratory). PRO

CONTIN analysis has been performed with a quiet data set.

To apply PRO CONTIN to the PBO/M SA polyelectrolyte system w e have

collected 50 to 200 runs and selected from these data files totaling more than 15,000 s

acquisition time. The correlation function was sufficiently quiet to perform the PRO-

CONTIN analysis. Figure 5.11 (a), (b), (c) and (d) show examples o f the PRO

CONTIN analysis o f high and low salt samples. One EXSAM P, which is a Laplace

inversion algorithm [5.26], fitting plot is also in Figure 5 .11(e).

142

PRO COTIN 50° Vv High Salt0 .0 8 -

0 .0 6 -

0 .0 4 -<

0 .0 2 -Avg.=4.742+/-0.514

2.713+/-0.3160 .0 0 -

651 2 3 4 7 8

log (T)

Figure 5 .11(a) PRO CONTIN analysis o f low molecular weight PBO/M SA-M SAA solution at high salt condition: M = 19,000; Cone. = 4 x l0 ' 3 g/ml; cs = 0.45 M; Vv measurement; 50° scattering angle; 514.5 nm wavelength; 600/100 pin-hole setting; ~15,000 s acquisition time with multiple runs; 25°C. The fast mode decay is broad in this plot.

143

0 .0 8 -□ PRO CONTIN 60° Vv

High Salt

□$

0 .0 6 -

< 0 .0 4 -

0.0 2 -

0.0 0 -

$

*

□□

j p Ayg

Avg.= Avg.=4.438+/-0.239 5.388+/-0.250

2.857+/-0.341

0 0 ° raP a m

1------r— r2 3

1------'------T4 5

log (T)

T— '— r

Figure 5 .11(b) PRO CONTIN analysis o f low molecular weight PBO/M SA-M SAA solution at high salt condition: M = 19,000; Cone. = 4 x l0 ' 3 g/ml; cs = 0.45 M; Vv measurement; 60° scattering angle; 514.5 nm wavelength; 600/100 pin-hole setting; ~ 15,000 s acquisition time with multiple runs; 25°C. The fast mode decay appear to consist o f tw o sub-modes.

144

0.020 -

0 .0 1 5 -

< 0 .0 1 0 -

0.005 -

0 . 0 0 0 -

□ PRO CONTIN 60° Vv,, Low Salt

A0 □0 □

0 □□ D

•o-

□ -o-

□ ^

□ ^

□ 1̂

□ CT]_L

D 6

D 6□ ' h

□□

0□

0□ □

□ Avg. = n □

4.318+/-0.861 \

12

i3 4

log (r)

T5

Figure 5 .11(c) PRO CONTIN analysis o f low molecular weight PBO/M SA-M SAA solution at low salt condition: M = 19,000; Cone. = 4 x l0 ‘ 3 g/ml; cs = 0 M ; Vv measurement; 60° scattering angle; 514.5 nm wavelength; 800/200 pin-hole setting;~ 15,000 s acquisition time with multiple runs; 25°C. A broad fast mode decay is in the plot.

145

□ PRO CONTIN 60° Hv HjHigh Salt

0 .0 1 5 -

0 .0 1 0 -

0.005 -

Avg.=5.046+/-0.643

0 . 0 0 0 -

1 2 5 7 83 4 6

log (r)

Figure 5 .11(d) PRO CONTIN analysis o f depolarized scattering for low molecular weight PBO/M SA-M SAA solution at high salt condition: M = 19,000; Cone. = 4x10 ' 3

g/ml; cs = 0 M; H v measurement; 60° scattering angle; 514,5 nm wavelength; 800/200 pin-hole setting; ~ 15,000 s acquisition time with multiple runs; 25°C. A broad fast mode decay is in the plot.

146

□ EXSAMP Fitting0 . 4 -

0 . 3 -

Avg. = 5.50+/-0.25Avg. =

4.50+/-0.30

I Avg. = P C □ip 2.86+/-0.34 ^0 .0 -

1 2 3 4 6 85 7

log(r)

Figure 5 .11(e) EXSAM P analysis o f low molecular weight PBO/M SA-M SAA solution at high salt condition: M = 19,000; Cone. = 4 x l0 ‘ 3 g/ml; cs = 0.45 M; Vv measurement; 60° scattering angle; 514.5 nm wavelength; 600/100 pin-hole setting;~ 15,000 s acquisition time with multiple runs; 25°C. The fast mode decay appear to consist o f tw o sub-modes.

147

M ost o f the correlation functions show a single broad peak in this fast decay region.

The PRO CONTIN and EXSAM P distributions could not clearly resolve the fast decay

mode, even though some o f them suggest two sub-modes (Figure 5 .11(b)). The decay

rate o f the faster mode, log(F) = 5.4, in this two-sub modes is similar to the decay o f

the H v measurement, lo g (0 = 5.0. The slow mode in this sub-modes is the same decay

range as the H v decay at low salt, log(T) = 4.5.

The amplitude o f the fast mode at low salt is comparable for the amplitude o f

the fast mode at high salt. It seems that the fast mode o f high salt doesn't get disrupted

by the dominant slow mode.

The slower mode in the Vv measurement o f high salt condition shows linear

angular dependence. The T vs. q2 plot gives 2 .68x l0 " 8 cm2/s (Figure 5.12), and it

indicates the translational diffusion coefficient o f the rod-like polymer. Using the three

initial angles, one obtains D t = 2.33x1 O' 8 cm2 /s, and the intercept at q = 0 gives Dt =

2.16x 10~8 cm 2 /s. The T/q 2 vs. q2 plot also shows the clear angular dependence o f the

slow mode. The measured diffusion coefficients are a little faster than the expected

translational diffusion coefficients o f the same polymer dimensions.

The fast mode decay is very sensitive to selection o f channel number and it is

hard to get the quantitative diffusion coefficient. The scale o f the fast mode is shown

for different experimental geometries in Figure 5.13. The fast mode in both Vv and H v

geometries at high salt is faster than that o f low salt, which means the fast mode o f high

salt is coupled with the dominant slow diffusion o f the molecules. This interaction

increasingly effects the fast mode decay. Experimental results shows that the fast mode

is more coupled in Vv geometry, which makes the fast mode slower than the H v. The

fast mode in low salt could be interpreted as non-coupled fast mode, T ~ 2 0 ,0 0 0 /s, in

this measurement.

148

CN

C/3

00

O(N

a -

4□ Slow Mode CONTIN

3

intercept = 2.16 x 10'

2

1

q2/ 1010 cm'2

2 -

■C/3

<3oT“"-<

1 -

0

0 2 4 6 8 10

q2/1 0 10 cm-2

Figure 5.12 Angular dependence o f the slow decay mode o f low molecular weight PBO/M SA-M SAA at high salt condition. M = 19,000; Cone. = 4 x l0 ' 3 g/ml; cs = 0.45 M; Vv measurement; 6 different scattering angles; 514.5 nm wavelength; 600/100 pin­hole settings; ~ 15,000 s acquisition time with multiple runs for each point; 25°C.

slope = 2.68 x 10'

slope = 2.33 x 10"

149

□ High salt fast Vv O High salt Hv

Low salt Vv

10 -

CO

'bu

1 . 1 1 1 1 1 1 1 1--------

0 2 4 6 8 10

q2/1 0 10 cm'2

Figure 5.13 Angular dependence o f the fast decay mode for low molecular weight PBO/MSA-M SAA. M = 19,000; Cone. = 4 x l0 ' 3 g/ml; cs = 0.45 M and 0 M; Vv and H v measurement; 4-5 different scattering angles; 514.5 nm wavelength; 600/100 pin-hole settings for Vv high salt, 800/200 for Vv low salt and H v; ~ 15,000 s acquisition time with multiple runs for each point; 25°C.

150

5.3.4 DISCUSSION

The translation and rotational diffusion o f the rod-like polymer can be

calculated by several means. The diffusion coefficient in the Kirkwood-Riseman (ICR)

approach for the rod-like polymer is

where D°t and D°r are the translational and rotational diffusion coefficient at zero

concentration, D° = kT/f°, where f° is the zero concentration molecule friction

coefficient, L is the contour length o f polymer, d is the rod diameter, and r | 0 is the

solvent viscosity. The calculated translational and rotational diffusion for L = 98 and

400 nm PBO are summarized in Table 5.3. Low molecular weight PBO shows

~ l x l 0 " 8 cm2/s for the translational diffusion coefficient and ~ 1 ,0 0 0 /s for the rotational

diffusion. W hen we consider two exponentials, from the equation T j = q2 Dt and r 2 =

q2 Dt + 6 Dr, the corresponding decay rates for translational and rotational motion are ~

1,000 and 5,000/s, respectively.

The translational diffusion coefficient obtained from experiments at high salt

condition (0.45 M ) agrees reasonably well with the theoretical KR approach. When we

use the data set o f Table 5.2 and extrapolate to zero concentration, the zero

concentration diffusion coefficient, D°t , is 1.32xl0"8 cm 2 /s. This value is around 30%

different from the K R approach. The measured fast decay, however, is much faster

than the decay associated with the rotational diffusion coefficient.

When we consider the polyelectrolyte's effect and charge interaction o f the

polymer, the electrostatic contribution on the polymer dimension can be obtained from

(5.1)3701^

3kT ln(—)(5.2)

151

Eq. (3.11). The electrostatic diameter o f charged rod, dH, is 4 - 50 times bigger than

the bare rod diameter, d0, by decreasing the ionic strength. The ionic strength o f 0 M

salt solution is approximated from the dissociation constant o f the solution [5.25],

Definition o f each length scale is in Chapter 1. Long range intermolecular repulsion

effects increase the translational diffusion in this case.

Table 5.3. The calculated translational and rotational diffusion from the Kirkwood- Riseman approach. r | 0 is 14.89, 15.03, 16.29, and 22.53 cP for 0 M, 0.01 M, 0.10 M, and 0.45 M salt solution, respectively. It is assumed that d = 0.5 nm. The unit o f translational diffusion is 1 0 ‘ 8 cm2 s '!and that o f rotational diffusion is s '1.

N o Salt 0.01 M 0.10 M 0.45 M

98 nm

400 nm

D°t 1.58 1.56 1.44 1.04

F ° t 1070 1060 975 705

D°t .490 .485 .448 .324

98 nm

400 nm

D°r 1480 1460 1350 980

F ° r 8880 8760 8100 5990

D°r 27.5 27.3 25.2 18.2

Then we consider the number density, v, o f the solution, the number density o f

4 x l0 ' 3 g/ml solution is 12 .7x l0 1 6 /cm 3 for M = 19,000 and 3 .09x l016/cm3 for M =

78,000, respectively. The calculated critical concentration, v* = 16/7rdHL2, where the

98 nm rods (M = 19,000) in 0.45 M salt condition (dH = 1.71 nm) start to overlap is

20 .7 x l0 1 6 /cm 3 and for 400 nm rod (M = 78,000) is 1 .76xl016/cm3. The critical

concentration at 0 M salt condition is 1 .24xl016/cm3 for 98 nm rods, which means

around 10 (v/v* = 12 .7xl016/1 .2 4 x l0 16) rods are interacting in the L 3 space. This

approximation provides neighboring interaction between the charged surface o f the

rods and the possibility o f the repulsion o f the rods. By decreasing the salt

152

concentration, the increased electrostatic diameter initiates the interaction between the

rods.

Table 5.4 The dimension o f PBO from Odijk's polyelectrolytes consideration.

OM 0.01 M 0.10 M 0.45 M

Ionic Strength / ( M) 0.0035 0 . 0 1 0 . 1 0 0.45

Bjerrum Length ■S„(nm) 0.56

Debye Length k "1 (nm) 5.81 3.44 1.09 0.51

Charge Distance b (nm) 0 . 6

Electrostatic Diameter d„ (nm) 28.4 15.23 3.92 1.71

Bare Rod Diameter d0 (nm) 0.5

The correlation functions at high salt condition which have tw o decay modes

suggest the polydispersity o f sensible dimensions. For example, it could be interpreted

as arising from some small oligomeric molecules in low salt condition. However, the

fast mode in depolarized measurement does not correspond to the translational

diffusion. This issue is very controversial in polyelectrolytes phenomena. Another

possibility to explain these tw o mode decays is coupling motion o f the polyelectrolytes.

When repulsive interactions are strong, it is dominant at low salt condition which gives

bigger electrical double layer, solutions exhibit very small fluctuations away from their

equilibrium positions. This condition, called an osmotically stiff solution, gives weak

scattered light and rapid decaying correlation function. As salt is added, larger

fluctuations are allowed leading to greater scattering.

The fast mode decay is hard to understand, but it is not an instrumental artifact.

Every experiment has been performed under high enough scattering intensity to prevent

instrumental effects in the very fast signal. The low intensity polystyrene/toluene

153

system has a nice correlation function but it also has a low f(A) value. Even though the

fast mode could be a tail o f the even faster decay o f the correlation function, the LS

instrument can not reach that fast decay. This is not evidence o f the smaller particles'

motion because it doesn't have angular dependent behavior. These experimental results

suggest the fast decay mode does not trace the polymer translational mode, but the

rotating, bending or coupling motion o f the counterion surrounding the

polyelectrolytes.

5.4 SUMMARY

The synthetic rod-like polyelectrolytes show the normal polyelectrolyte

behavior, high coherence in high salt and low intensity at low salt condition. O-E

transition which has a very slow mode in low salt condition is one o f the interesting

features o f polyelectrolyte solutions, but we couldn't see any very slow decay ( 1 0 times

bigger than the normal polymer diffusion) motion except the dust interruption o f the

correlation function. The slow decay mode in this experiment shows the diffusion o f

the polymer backbone in the high salt condition. The fast mode is still hard to

understand, but it is not the translational contribution o f particle diffusion.

154

5.5 REFERENCES

5.1 (a) L. Belloni and M. Drifford, J. Phys. Lett. 46, LI 183 (1985).

(b) W. I. Lee and J. M. Schurr, J. Polym. Sci. 13, 873 (1975).

5.2 C. F. Anderson and M. T. Record Jr., Ann. Rev. Phys. Chem. 33, 191 (1982).

5.3 M. M -Noyola and A. V-Rendon, Phys. Rev. A. 32, 3596 (1985).

5.4 T. Odijk, Macromolecules 12, 688 (1979).

5.5 M. Drifford, L. Dalbiez, and A. K. Chattopadhya, J. Coll. Inter. Sci. 105, 587 (1985).

5.6 N. Ise, T. Okubo, K. Yamamoto, H. M atsuoka, H. Kawai, T. Hashimoto, and M. Fujimura, J. Chem. Phys. 78, 541 (1983).

5.7 B. R. Ware, Adv. Coll. Inter. Sci. 4, 1 (1974).

5.8 D. F. Hodgson and E. J. Amis, J. Chem. Phys. 91, 2635 (1989).

5.9 M. Levij, J. de Bleiser, and M. Leyte, Chem. Phys. Lett. 83, 183 (1981).

5.10 H. M atsuoka, N. Ise, T. Okubo, S. Kunugi, H. Tomilama, and Y. Yoshikawa, J. Chem. Phys. 83, 378 (1985).

5.11 N. Ise, T. Okubo, S. Kunugi, H. M atsuoka, K. Yamamoto, and Y. Ishii, J. Chem. Phys. 81, 3294 (1984).

5.12 S. -C. Lin and J. M. Shurr, Biopolymer 17, 425 (1978).

5.13 M. Drifford and J. -P. Dalbiez, Biopolymer 24, 1501 (1985).

5.14 M. Drifford and J. -P. Dalbiez, J. Physique Lett. 46, L 311 (1985).

5.15 K. S. Schmitz, L. Mei, and J. Gauntt, J. Chem. Phys. 78, 5059 (1983).

5.16 G. C. Berry, Contemp. Top. in Polym. Sci. 2, 55 (1977).

5.17 M. A. Tracy and R. Pecora, Annu. Rev. Phys. Chem. 43,525 (1992).

5.18 D. B. Roitman, J, McAlister, M. McAdon, E. M artin and R. A. Wessling, J. Polym. Sci. Polym. Phys. 32, 1157 (1994).

5.19 C.-P. W ong, H. Ohuma, and G. C. Berry, J. Poly. Sci., Poly. Sym. 65, 173 (1978).

5.20 K. Schmitz, "An Introduction to Dynamic Light Scattering by M acromolecules," Academic Press Inc., 1990.

155

5.21 P. M. Cotts and G. C. Berry, J. Polym. Sci., Polym. Phys. Ed. 21, 1255 (1983).

5.22 M. Tracy, Ph. D. Thesis in Stanford University, 1992.

5.23 C. R. Crosby, III, N. C. Ford, Jr., F. E. Karasz, and K. H. Langley, J. Chem. Phys. 75(9), 4298 (1981).

5.24 W. Eimer and R. Pecora, J. Chem. Phys. 94, 2324 (1991).

5.25 D. B. Roitman, R. A. Wessling, and J. McAlister, Macromolecules 26, 5174 (1993).

5.26 P. Russo, K. Guo, and L. DeLong, Soc. Plastics Eng., 46th Ann. Tech. Conf. Pro. 983 (1088).

CHAPTER 6

CONCLUSIONS

156

157

There are two aspects in this dissertation, one is about the probe particles and

the other is about the matrix polymers. Probe particles reveal superparamagnetic

properties and charged interaction, and the matrix polymer (polystyrene sulfonate

sodium salt: NaPSS) shows charged polyelectrolyte behavior. Chapter 2 and 3 are

concerned with the interaction between magnetic latex particles and the matrix polymer

under different salt conditions. One o f the issues was flocculation and stabilization

interactions between the particle and matrix polymer. Dynamic light scattering and

fluorescence photobleaching recovery have been used to trace the diffusion o f MLP

and NaPSS. A DLS study with varying salt concentrations shows the flocculation o f

the particles in the high salt condition and the stabilization o f MLP by adding matrix

polymer. DLVO theory explains the flocculation phenomena by decreasing the Debye

screening length in high salt and by increasing the attractive force between the particles.

The stabilization mechanism for the particles and matrix polymer has been suggested by

Feigin and Napper, but little experimental evidence has been reported. FPR

experimental results using labeled NaPSS suggested that the particles and matrix

polymers depleted each other and stabilized the particles' flocculation (depletion

stabilization).

A magnetic field induced study using superparamagnetic properties o f the probe

particles has been performed and presented in Chapter 4. When the magnetic field is

applied, superparamagnetic particles form linear chains (columns) parallel to the applied

field. The applied field and concentration dependence o f the magnetic latex particles'

kinetic growth was traced by microscopy and video small angle light scattering.. The

high magnetic susceptibility o f the MLP explains why they align at very weak magnetic

fields. A scaling concept is applied for this cluster size growth mechanism and the rate

o f cluster formation is compared with SALS and MS techniques. Good agreement was

158

obtained between the tw o methods, and s(t) ~ t 1 0 was observed in the initial growth o f

the cluster.

The polyelectrolytes studies were extended in Chapter 5. "O-E transition" and

"ordering" structure o f polyelectrolyte solutions which are controversial topics related

to polyelectrolyte solutions are considered to understand the DLS data o f a high

strength rod-like polyelectrolyte (PBO). The PBO/M SA-M SAA system shows the

normal polyelectrolyte behavior, high coherence in high salt and low intensity at low

salt condition. However, it does not show any very slow decay in DLS measurement

( 1 0 times bigger than the normal polymer diffusion) which has been typically observed

in several polyelectrolyte solutions. There are tw o decay modes in our DLS

experiment. The slower one presents the diffusion o f the polymer backbone in the high

salt condition and the fast mode is not related to. This is explained by the expanded

concept o f the osmotic stiffness.

The motion o f the magnetic particles in a chemically cross-linked gel has been

studied and summarized in Appendix A. It is a preliminary study o f the morphology o f

silica gel and acrylamide gel, and their affinity with the magnetic latex particles and

ordinary latex particles. It should lead to a future study to explain the magnetic

particles' aggregation in silica gel, to perform the angular dependence o f the probe

particles in the gel, and to trap the stacked magnetic particles' in the gel structure.

APPENDIX A

THE BEHAVIOR OF MAGNETIC LATEX PARTICLES IN CHEMICALLY CROSS-LINKED GELS

159

160

A.l INTRODUCTION

A gel is a fluid-filled cross-linked polymer network. Depending on the methods

o f cross linkage o f the polymer, chemical and physical gels are classified. The structure

o f chemically and physically cross-linked gels has been studied by techniques such as

cold-stage electron microscopy, video-optical microscopy, etc. [A.1] Dynamics such

as diffusion in or viscoelasticity o f the gel network have been extensively studied by

light scattering [A.2] and rheo-optical techniques [A.3],

There are tw o approaches to the DLS technique as applied to gel dynamics; one

is the viscoelastic approach based on Tanaka's 1973 paper [A.4], and the other is the

non-ergodic approach o f Pusey et al.[A.5]

In 1973, Tanaka et al. studied polyacrylamide gels and proposed that the

autocorrelation function o f scattered light was due to thermally excited density

fluctuations in fiber-network. The gel was considered to be a uniform continuous

medium that undergoes spontaneous thermal fluctuation, modifies the dielectric

permittivity, and hence scatters light [A. 6 ], The thermally excited displacement o f the

gel network with ( t t ) or against ( t 4 ) the gel liquid are characterized by the (a)

elastic constant o f the fiber network, and (b) the friction factor connecting the gel

network and the gel liquid. They derive the time correlation function o f the scattered

light from the gel and give the decay constant T = 1/t =

where K' and p' are the bulk and shear moduli, respectively, q is the wave vector o f the

fluctuation and f is the frictional constant. The macroscopic measurement o f the

frictional (f) and elastic constant (K1, p') were within the 1% experimental error o f the

dynamic light scattering analysis. M ost DLS studies with gels had been based on this

polarized (A.1)

( 2 p '/ f ) q 2 : depolarized (A.2)

161

viscoelastic continuum picture o f the gel network until Pusey et al.' s non-ergodic

framework was developed.

In a DLS experiment the normalized time correlation function g<2 )(q,x) o f the

scattered intensity is given by

g(2 )(q ,x ) = < I(q ,0 )I(q ,T ))/ ( l(q ,0 ) ) 2 (A.3)

= l + f(A )|g (1)(q ,x ) | 2

The angular brackets imply an ensemble average, but it can be replaced with a time

average based on the ergodicity theorem in most DLS experiments. In a gel the total

scattered field may be written [A. 5] as the sum o f the fluctuating component Ep(q,t)

and a time independent component E c (q)

E(q,t) = EF(q,t) + E c (q) (A.4)

The time-average o f the total scattered intensity can be expressed as

( l ( q » T = ( |E (q ,x |2)T

= (IF(q ))T + I c (q) (A .5 )

Time-average intensity correlation function is a mixture o f a Gaussian and a constant

field, and leads to

(l(q> 0 )I(q , x))T - ( l(q , 0 ) ) 2

= (i(q))E [g(NE(q»'c) -g N E (q .Q0) f ,(A. 6 )

+2Ic (q )( l(q ))E[g & (q ,x )-g $ s (q ,o o )]

The first term on the right side is the homodyne contribution and the second term is the

heterodyne part o f the total intensity correlation. The mixed homodyne-heterodyne

case is treated by Chu [A.7] and previously applied in our lab (details in [A.8 ]).

The characterization o f the probe particles in polymer gels gives very unique

information about the structure and morphology o f the gel. The diffusion o f the probe

162

particles in a polymer gel depends on the diameter o f the probe, d, and the correlation

length, £ o f the network. I f d » £ we can predict that since the particles are bigger

than the gel structure, there is no long-range motion o f the particles. I f the particles are

smaller than the gel network, d « £ there is finite diffusion o f the particles. The

diffusion o f the particles inside o f the solid cage is a very unusual experiment. Drifford

et al. [A. 9] did the diffusion studies with calibrated particles during the formation o f a

gel, and the mesh size £ at the end o f the chemical reaction is lower than d. Konak et

al. [A. 10] reported a QELS study o f polystyrene latex particles in polyacrylamide gel

to determine the static gel structure. These studies predate the non-ergodic

developments. Joosten et al. [A. 11] investigated the Brownian particles trapping in

polyacrylamide gel and used non-ergodic concepts.

Some polymer gels themselves give a strong polarized signal (e.g. PAA gel),

but their depolarized signals are almost negligible. Therefore, the polarized signal o f a

probe particle inside the gel is overlaid by the scattering from the gel network, but the

depolarized signal is independent. Here we investigate tw o different types o f chemical

gels. PAA gels have pronounced viscoelastic properties and produces a strong

polarized signal, silica gels have plastic-like properties and do not produce any

polarized signal. And tw o different particles, normal latex and magnetic latex particles,

are used as a probe for the translational and rotational studies.

A.2 EXPERIMENTAL

A.2.1 LATEX PARTICLES

The probe particles used in this experiment are small (diameter ~ 0.05 pm)

polystyrene latex and magnetic latex particles that are commercially available (Details

o f the magnetic latex particles are in Chapter 2 and 3.) Each type o f particle was

purchased from Polysciences or Bangs Laboratories as 2.5 w t% o f solids solutions.

Polysciences particles are based on styrene monomer with a sulfonated charge on the

163

surface, and Bangs particles are carboxylated charged. Magnetic contents o f the MLP

is w 10% o f the total latex weight. It creates a weak depolarized signal, but enough to

measure. Bangs magnetic latex particles are filtered with 0 . 1 pm Millipore filter before

mixing with the gel precursor solution. Each particle is well dispersed in the water

solution by the surfactant, and gives a single exponential decay.

A.2.2 SILIC A G E L

Silica gels are the precursors o f aerogels [A. 12] which are low density porous

solids dried without shrinking. The no-shrinking condition can be achieved by the

hypercritical extraction or drying in which no surface tension exists between the liquid

surface and the air. So far, porosities o f 99.8% (0.2% polymer) have been reported

and many researcher's are involved in finding applications for these materials. The

most popular applications are as insulators due to their high heat capacity, and as

catalysts for organic synthesis.

The silica gel preparation consists o f 2 steps: a) preparation o f a sol-gel stock

solution by acid catalyzed esterification; and, b) gelation by basic catalyzed and aging

(polycondensation). To make a stock solution, 30 ml TEOS, 30 ml ethanol, and 2.5 ml

deionized water were mixed in a round bottom flask. As a catalyst, 0.1 ml o f 1 M HC1

was added. The molar ratios among the TEOS, EtOH, and water were 1:4: 1. The

mixed solution was stirred for 1.5 hours at 60°C. Then the sample was cooled down to

room temperature. This was the precursor (stock solution) for the silica gel. In the

next step, 0.2 ml 0.1 M NH4OH was added to the 2 ml o f stock solution, followed by

filtration with 0.1 pm Millipore M illex® -W filter, in a cylindrical light scattering cell.

The gelation time was varied by the ratio o f TEOS/EtOH/H20 and mostly by the

amount o f the added base. Typical gelation time was 15 min. to 3 hrs. Different molar

ratios o f TEOS/EtOH (1/80, 1/40, 1/33, 1/27, 1/20, 1/10, 1/5, 1/4) have been made by

164

the same method. Normal o r magnetic latex particles 0.05 ml are added to the pregel

mixture prior to the initiation o f the gelation by ammonium hydroxide catalyst.

A.2.3 ACRYLAMIDE GEL

Acrylamide gel is well known because o f its biochemical applications, especially

in electrophoretic separations [A. 13]. Polyacrylamide gels or solutions are made by

copolymerizing acrylamide (AA) and N, N-methylene bisacrylamide (BAA) using

ammonium persulfate (AP) as an initiator and N, N, N*, N’-tetramethylethylenediamine

(TEM ED) as the activator. In 10 ml o f clean water, O.Olg (6.48x1 O' 3 moles) BAA

(Aldrich) and 0.24 g (0.34 moles) AA (Aldrich) were dissolved. This was a stock

solution o f 2.5 wt% and 4% BAA/AA ratio. The various concentration (0.25 g, 0.30

g, 0.35 g, 0.45 g) and BAA/AA ratio (1% - 40%) stock solution were made by the

same method. The stock solution was dedusted using 0.1 pm Milipore W type filter.

Stock solution 1 ml was put into the LS cell, and added 2.5% 0.05 ml normal or 0.2%

0.05 ml magnetic latex particles. Final ordinary latex and M LP concentrations are

0.13% and 0.01%, respectively. TEM ED (Aldrich, 99% ) 3 pi and 5 pi o f 0.5 M

ammonium persulfate were added to initiate the gelation. The cells was flushed with

argon and sealed.

A.2.4 M E A SU R E M E N T

The light scattering instrument was described in previous chapter. It consists o f

a He-Ne laser o f 632.8 nm for the polarized measurements and ion laser (514.5 nm) for

the depolarized experiments, focusing lens, vertical polarizer, the sample holder which

is connected with the ro tor to change the position o f the speckle o f the non-ergodic

samples, horizontal or vertical analyzer, and receiving lens, pin-hole, and EM I 9863

photomultiplier tube. An ALV-5000 digital autocorrelator is used to measure the

correlation function with multiple runs. Temperature during the measurement was

maintained at 25±0.1°C.

165

A.3 RESULTS AND DISCUSSION

A.3.1 ORDINARY LATEX PARTICLES IN SILICA GEL

As the goal o f this study is to m onitor the probe particles in the silica gel, we

chose the fast gelation process. The gelation (sol-gel transition) process o f the silica

gel is a good example o f the structure growth o f a fractal [A. 14], To monitor the

kinetic growth o f the silica cluster, DLS is used to determine the diffusion coefficient

and the hydrodynamic radius. The gelation process is relatively fast, tge] = 10 min. to 3

hr s.

Figure A. 1 shows an example o f the gelation o f the silica gel used in this

experiment, time dependent D app. Our gelation process is much faster than that o f the

diffusion limited aggregation o f Martin et. al. [A. 15] Since we are concerned with the

motion o f the particles in the gel network, we chose the fast gelation process. After the

gelation process is completed, the viscosity o f the gel rises abruptly and elastic

response to stress appears. The viscosity o f inorganic gels was reported ~ 101 2 Pa s

and the elastic shear modulus is « 107 Pa [A. 16], This is not viscoelastic behavior, but

near perfect plasticity or linear hardening. DLS experiments show no correlation after

the gelation process is completed. When the polystyrene latex particles are in the sol

state o f the silica precursor, the scattered intensity o f the latex particles is much larger

than that o f the growing clusters. There is a competition between the latex particles

and growing clusters, so it is too hard to separate the decay rate o f each different

species during the gelation process.

However, the latex particles' decay is evident in the completed gel structure

where there is no correlation o f the gel network itself. It is hard to compare the

intensity ratio between the latex particles and the gel network because o f the non-

ergodic properties, but the averaged ratio Iiatex in gei/Igei ~ 3.

166

6 0 -

5 0 -

CL,

Q 2 0 -

1 0 -

0 50 100 150 250 300200

Time (min)

Figure A. 1 Diffusion coefficient change o f the silica sol-gel transition with time. Sample composition was 0.5 ml TEOS, 10 ml EtOH, catalyzed with NH 4 OH. Light scattering measurement was done at 25°C, 90° scattering angle. Diffusion coefficient is from the 3rd cumulant o f the ALV correlator, (3 - 128 channels).

167

The diffusion coefficient o f the latex particles is 3.0x1 O' 8 cm2/s in the initial sol state

and it turns to 2 .7x l0 " 8 cm2/s in the completed gel, 0.01 M ratio o f TEOS/EtOH.

Figure A.2 shows the correlation function o f the latex particles in the gel and compares

it to latex particles in water. Even though the diffusion is slowing down, D = 7 .74x l0 ‘ 8

cm2/s in water and D ~ 3.0x10" 8 cm2/s in the gel structure, the latex particles are freely

moving inside o f the pores which implies that mesh size is bigger than the particle

diameter. The experiment shows a single exponential decay in this case.

When we change the molar ratio o f the TEOS/EtOH (5%, 10%, 12%, 15%),

the diffusion o f the latex particles slows down linearly. The high molar ratio o f

TEOS/EtOH gel, > 20%, does not show particles diffusion where the length scale o f

the light scattering, 27t/q, is around 450 nm. The dilute solution (< 2% ) doesn't make a

gel even after 1 month. Figure A.3 shows the correlation function o f the 0.059 pm

latex particles in different TEOS/EtOH ratio o f completed silica gel. TEO S/EtO H ratio

are 0.5%, 10%, 12%, 15%, respectively. Each correlation function is averaged from

m ore than 50 different spots o f the gel by rotating the sample using a com puter driven

stepping motor. M easurement was performed with a 632.8 nm He-Ne laser at 90°

angle and 25°C. The diffusion coefficient in the water, D°, is 7 .74x l0 ' 8 cm2/s which is

higher than that o f the initial silica solution, and the Dapp/D 0 vs. molar ratio o f

TEO S/EtO H is in Figure A.4. Two length scales are assumed from this plot.

168

2.0 -m waterin the completed gel

1.8 -

1.6 -

1 .4-

1.2 -

io'10°

t / s

in the completed gel

- 2 - in water

40 2

x / 1 0 ' 3 s

Figure A.2 Correlation function o f polystyrene latex particles, diameter = 0.059 pm, in the silica gel network, compared to that in the water. The composition o f the gel is 0.5 ml TEOS, 10 ml EtOH, catalysed with NH 4 OH. Measurement was performed at 25°C, 90° scattering angle.

169

0.5 ml TEOS 1.0 ml TEOS 1.2 ml TEOS 1.5 ml TEOS

2.0

10°

t/ s

0

□ 0.5 ml TEOSo 1.0 ml TEOS A 1.2 ml TEOS v 1.5 ml TEOS

1

•2200 10 155

t / 1 0 '3 s

Figure A.3 Correlation function o f the polystyrene latex particles, diameter = 0.059 pm, in different TEOS/EtOH ratio o f completed silica gel, after 2 days. TEOS/EtOH ratio are 0.5%, 10%, 12%, 15%, respectively. Each correlation function is averaged from more than 50 different spot o f the gel. Measurement is performed with 632.8 nm He-Ne laser at 25°C, 90° scattering angle.

app

170

0.5

0.4

0.3

Slope=6.74o

0.2

0.00.04 0.05 0.060.030.02

M Ratio o f TEOS/EtOH

0.00 0.01

0 .3 7 pm 0 .0 6 p m

Figure A. 4 Diffusion coefficient o f latex particles in silica gel decreasing by increasing the TEOS/EtOH molar ratio. D° = 7.74x10 ‘ 8 cm 2 /s.

171

The average distance o f the particles in the solution, 0.37 pm, should be the length

scale at 0 molar ratio o f TEOS/EtOH, which is estimated with the relation v ^ 3 = 1

where v is number density in 2 ml solution at 2.5% and £ is correlation length. W e

assumed correlation length is the same as the diameter o f a particle, 0.059 pm, when

there is no motion o f the particles at 0.058 M ratio.There is a linear relation between

the the correlation length and the diffusion coefficient. The linear relation o f

correlation length in Figure A.4 means the gel structure is more uniformly shrink than

that o f the PAA gel. M ost o f the correlation signal is coming from the fluctuating part

o f the gel/particle ternary structure, not from the gel structure. Non-ergodic

considerations o f the ICF explains a small shift in the baseline, but no dramatic shift in

ftA). The shift o f the f(A) in this ternary system can tell the trapping o f the particles in

gel structure.

A.3.2 MAGNETIC LATEX PARTICLES IN SILICA GEL

0.05 pm magnetic latex particles in the silica gel did not move in the gel

structure, even in the soft gel (« 3% TEOS/EtOH). Figure A .5 shows the gelation

process o f the magnetic latex particles in the silica gel. As the gelation process

proceeds the particles are trapped in the gel structure. There is no correlation function

after completed the gel, after 24 hours. The particles aggregate when incorporated in

to the gel precursor. The aggregation can be seen with the naked eye at the high M LP

concentration. Probe diffusion studies with the M LP in silica gel are not suitable for

these reason. The aggregation process might be caused by the charged interaction o f

the particles in the silica sol precursor.

A.3.3 MAGNETIC LATEX PARTICLES IN PAA GEL

A.3.3.1 TRANSLATIONAL DIFFUSION

DLS with normal latex particles in the PAA gel has been studied by several

research groups [A. 17, A. 18],

172

2.40 time 5 hours 24 hours2.2

2.0

0.810'7 10"6 10’5 10-4 10'3 10'2 10'1 1 0 °

t / s

Figure A. 5 Autocorrelation function o f the magnetic latex particles during the silica sol-gel transition. There was no correlation function after 24 hours, which means the M LP are trapped in the gel. The gel condition was 0.5% TEOS and measurement was performed with 632.8 nm He-Ne laser, 600/50 pin-hole setting at 25°C, 90° scattering angle.

173

Their observations show a complex dynamical behavior ranging from the purely

translational diffusion o f latex particles in the medium to a relaxational (slowing

down) behavior associated with local movement o f probes in the gel. M ore detailed

consideration o f the PAA gel will be discussed later.

M agnetic latex particle doesn't show any serious aggregation in the PA A gel

structure, and their correlation function decays like that o f normal latex particles in the

gel. By increasing the (BAA+AA)/water ratio up to 5 wt% , 0.5 g/10 ml water, and the

ratio (BAA/AA) up to 20 %, PAA gel becomes too turbid to measure the scattered

intensity. Vv (actually Uv, but it is the same in all measurements) and H v measurement

are measured before the onset o f turbidity. The particle's autocorrelation function from

the polarized light scattering in 3 different concentration (2.5, 3.5, 4.5% PAA) o f PAA

gel (BAA/AA = 0.04) is shown in Figure A.6 . This is the sum o f at least 50 different

ICFs (intensity correlation functions) o f different speckle patterns. One feature is

obvious from this plot: f(A) is decreasing with increasing gel concentration (weight %

o f BAA+AA) which indicates partial heterodyning due to the emergence o f strongly

scattering and stationary clusters. During the sample time scale, 60 s, which is enough

to get a reasonably clean ICF, the correlation function is not shifted from the baseline.

Non-ergodic considerations using Eq (A. 7) are shown in the Figure A. 7. These two

plots make it clear that the trapped particles act as local oscillators which cause partial

heterodyning. By decreasing the correlation length, £, the correlation function shifts to

the low time scale in case o f the silica gel, but the f(A) and baseline shift in the

acrylamide gel. In the AA gel the finite value o f f(A), fluctuating component, is coming

from the diffusion o f the particles in the gel, and the baseline shift, stationary

component, is due to the gel heterodyning. In each case the ratio between the

fluctuating and stationary component is 0.89, 0.49, and 0.19 for 2.5%, 3.5%, and 4.5%

PAA gel, respectively.

174

1.8 -

1.6 -

CN 1 . 4 -bJ)

1.2 -

1.0 -

□ □ 2.5% PA A

A

//• ■'/// /// // /// /// ,M •/-/ ///. ///. /it .//' //• // . ,

■I i » u u q— i 11 iiibh ‘ ■ i i m m)— i 1 111114— 1 1 m m il)— 1 1 m i i q — r i 'iT in n — 1 1 m i l l ) 1

10‘7 10'6 10'5 10‘4 10'3 10'2 10’1 10° 101

t / s

Figure A. 6 The time averaged autocorrelation function from the 50 different spots o f the M LP/PAA gel system. The particles' autocorrelation function is coming from the polarized light scattering in 3 different concentration (2.5, 3.5, 4.5 wt% ) o f PAA gel. The measurement was performed with 632.8 nm He-Ne laser, 600/50 pin-hole setting at 25°C, 90° scattering angle.

175

2 .0 -

19%1.83

49%

CN 89% 1.47

2.5% PA A

3.5% PA A

4.5% PA A

1.09iSIffieiniiiiiiiBMuiiMuiittJHiiimuiBiiimnnfflr

x/sec

Figure A.7 The ensemble averaged autocorrelation function from the 50 different spots o f the M LP/PAA gel system. The particles' autocorrelation function is coming from the polarized light scattering in 3 different concentration (2.5, 3.5, 4.5 wt% ) o f PAA gel. The measurement was performed with 632.8 nm He-Ne laser, 600/50 pin-hole setting at 25°C, 90° scattering angle.

176

A.3.3.2 ROTATIONAL DIFFUSION

Even though scattering intensity o f the particles is much greater than that o f the

PAA gel, the acrylamide gel has a decaying ICF in these concentration range (« 4

wt% ). The polarized scattering from the gel originates from the thermal fluctuation in

the gel network. The depolarized scattering o f the PAA gel is small and no one has

observed the depolarized spectrum from these gels. So, depolarized scattering with

optically anisotropic particles could be a very powerful technique. In our lab zero

angle depolarized scattering have been done for the gel system using the geometrically

and optically anisotropic PTFE particles [A.8 ], Magnetic latex particles are the other

challenge for depolarized probe diffusion studies with the gel. Although the

commercially available small magnetic particles have relatively low magnetite contents,

11 wt% o f total latex weight, there is some depolarized signal. The ion laser (514.5

nm), ~ 750 mW maximum intensity, is used and the power o f the laser is carefully

adjusted to give the depolarized signal and not to give the thermal agitation o f the gel

structure («0.5 mW). The concentration o f the MLP is increased to give a usable

depolarized signal and to have no concentration drift within the small detecting volume.

The depolarized experiment is performed at a constant temperature o f 25° at a

scattering angle o f 30°. Figure A . 8 shows the measured ICF, represented by the

decaying curve o f the depolarized scattering with ergodic and non-ergodic summing

from the ALVAN. Correlation function has low f(A) and more than 3% o f PAA gel

shows baseline shift.

A .4 S U M M A R Y

Probe diffusion with magnetic latex particles in tw o different types o f

chemically cross linked gel, silica gel and acrylamide gel, has been studied. MLP are

stable in the PAA gel, but tend to aggregate during the gelation o f the silica gel.

177

1.3

Aro

1 . 2 -

<NW)

1.1 -

1.0 -

a 3.0% PAA v 3.5% PAA o 4.0% PAA

_ l i m■ "I ■ ■ ■ ill ■_■ " ’I ■ ■ ml ■ ■ i ill J I I I I ! I L JU 11_]__■ m l

10'7 10'6 10'5 10'4 10'3 10'2 10’1 10° 101 102

t/ s

Figure A . 8 The depolarized autocorrelation function from the 50 different spots o f the M LP/PAA gel system. The particles' autocorrelation function is coming from the depolarized light scattering in 3 different concentration (2.5, 3.5, 4.5 w t% ) o f PAA gel. The measurement was performed with 514.5 nm He-Ne laser, 800/200 pin-hole setting at 25°C, 30° scattering angle, non-ergodic sum.

178

The diffusion o f particles in inorganic silica gel is decreased by decreasing the gel

correlation length, but the particles' diffusion in the viscoelastic (acrylamide) gel still

shows the heterogeneity o f the gel. By increasing the contents o f the gel network, the

silica gel makes a m ore homogeneous and compact structure than that o f the PA A gel

which has partial heterodyning. QELS study with probe particles in tw o different gels

w as successful to reveal the heterogenity o f the gel network.

In addition to the translational diffusion o f the particle, M LP are used as probes

for the depolarized scattering experiment. Rotational motion in the gel also support the

data for translational diffusion with regard to non-ergodic concepts.

179

A.5 REFERENCES

A .l R. D. Allen, Scientific American 256(2), 42 (1987).

A.2 R. Pecora, "Dynamic Light Scattering," Pleum, NY, 1985.

A. 3 B. Chung and A. -E. Zachariades, "Reversible polymeric Gels and Related Systems," P. S. Russo Ed., p22, 1986.

A.4 T. Tanaka, L. O. Hocker, and G. B. Benedek, J. Chem. Phys. 59, 5151 (1973).

A.5 P. N. Pusey and W. van Megen, PhysicaA. 157, 705 (1989).

A . 6 E. Geissler, "Dynamic Light Scattering," W. Brown, Chapter 11, p470, 1993.

A.7 B. Chu, "Laser Light Scattering," Academic Press, N ew York, 1991.

A . 8 B. Camins and P. S. Russo, Langmuir, in press.

A.9 C. Allain, M. Drifford, and B. G.-Manuel, Polymer Communication 27, 177 (1986).

A. 10 C. Konak, R. Bansil, and J. C. Reina, Polymer 31, 2333 (1990).

A l l J. G. H. Joosten, E. T. F. Gelade, P. N. Pusey, Phys. Rev. A. 42(4), 2161 (1990).

A. 12 Aerogel

(1) General

(a) C. J. Brinker and G. W. Scherer, "Sol-Gel Science," Academic Press, 1990.

(b) M RS Bulletin, XV(12), Dec. 1990.

(c) H. D. Gesser and C. Goswami, Chem. Rev. 89, 765 (1989).

(d) S. J. Teichnex, Chem tech, 372, June (1991).

(2) Preparation

(a) T. M. Tillotson and L. W. Hrubesh, J. Non-Cryst. Solids 145, 44 (1992).

(b) J. Phalippou, T. Woignier, and M. Prassas, J. Mat. Sci. 25, 3111 (1990).

(c) G. W. Scherer, J. Non-Cryst. Solids 147, 365 (1992).

(3) Properties and Applications

(a) A. M. Buckley and M. Greenblatt, J. Non-Crystals Solids 143, 1 (1992).

(b) D. W. Schaefer, B. J. Oliver, C. S. Ashley, D. Richter, B. Farago, B. Frick, L. Hrubesh, M. J. van Bommel, G. Long, and S. Krueger, J. Non-Cryst. Solids 145, 105 (1992).

(c) J. Fricke, J. Non-Cryst. Solids 147, 356 (1992).

(d) G. Poelz, "Preparation o f Silica Aerogel for Cerenkov Counter," DESY 81-055 DESY, Hamburg, 1981.

A. 13 J. Baselga, I. H-. Fuentes, I. F. Pierola, M. A. Llorente, Macromolecules 20, 3060 (1987).

A. 14 F. Family and D. P. Landau, "Kinetics o f Aggregation and Gelation," North- Holland, GA, USA, 1984.

A. 15 (a) J. E. Martin and J. P. Wilcoxon, Phys. Rev. A 39(1), 252 (1988).

(b) J. E. M artin and J. P. Wilcoxon, Phys. Rev. Lett. 61(3), 373 (1988).

A. 16 T. Tanaka, Encycropidia o f Polymer Science and Engineering, 2nd Ed. H. F. M ark et al. Ed., Wiley, New York, Vol. 7, p514, 1986.

A. 17 (a) I. Nishio, J. C. Reina, R. Bansil, Phys. Rev. Lett. 59, 684 (1987).

(b) S. Pajevic, R. Bansil, and C. Konak, Macromolecules 26, 305 (1993).

(c) J. C. Reina, R. Bansil, C. Konak, Polymer 31, 1038 (1990).

A. 18 L. Fang and W. Brown, Macromolecules 25, 6897 (1992).

APPENDIX B: UNITS FOR MAGNETIC PROPERTIES

181

182

Units for M agnetic Properties

Quality Symbol cgs emu ConversionFactor

SI & mks

Magnetic flux density, magnetic induction

B gauss(G) 10-4 tesla (T), Wb/m2

Magnetic flux <t> maxwell (Mx), G cm

10'8 weber (Wb), volt sec

Magnetic potential difference, magnetoactive force

U ,F gilbert (Gb) 10/47t ampere (A)

Magnetic field strength, magnetizing force

H oested (Oe), Gb/cm

103/47t A/m

Magnetization M emu/cm 102 A/m

Magnetic moment m emu, erg/G 103 A m2, J/Tesla

Magnetic dipole moment j emu, erg/G 4tcX 10 '10 Wb m

(Volume) susceptibility Xv Kv dimensionless,

emu/cm

471

(47c)2 x 10'7

dimensionless

H/m, Wb/A m

(Mass) susceptibility Xm’ Km cm3/g, emu/g 4tcX 10'3

(4tc)2 X 10‘10

m3/kg

H m2/kg

Permeability P dimensionless 4tcX 10'7 H/m, Wb/(A.m)

(Volume) energy density, energy product

W erg/cm 10"1 J/m3

Demagnetization factor D ,N dimensionless 1/471 dimensionless

* It is provided by Bangs Laboratoiy, Inc.

a. Gaussian units and cgs emu are the same for magnetic properties. The defining relation is B = H+AnM

b. Multiply a number in Gaussian units by C to convert is SI units.

APPENDIX C: COPYRIGHT PERMISSION

183

184

L S UL o u i s i a n a S t a t e U n i v e r s i t yA N D A G R I C U L T U R A L A N O M C C H A N l C

D e p a r t m e n t o f C h e m i s t r yC O L L l C I

June 7, 1994.

Ms. Janice Dininny Materials Research Society 9800 McKnight Road Pittsburgh, Pennsylvania 15237

Dear Ms. Dininny:

I am writing to you in reference to the article entitled, "Light Scattering from Magnetic Latex Particles," published in the Material Research Society Symposium Proceedings, Vol. 248, pp. 247-252 (1992), for which I am the first author. I would like to use the manuscript in my Ph. D. dissertation. Please forward permission for reprint of

the manuscript.

I will appreciate your prompt reply.7 / 3 0 / 9 4Permission granted. Please cite source.

mjce DininnyJarPublications Dept. Materials Research Society

Prof. Paul S. Russo

Sincerely,

odlcU Vtktn.

Daewon Sohn

VITA

The author, Daewon Sohn, was bom in Seoul, Korea on September 12, 1961

(by lunar calendar). He received his bachelor degree in chemistry from Hanyang

University in Korea in 1984. He obtained a master degree in chemistry in 1986 with a

thesis, "A New Flow Equation for the Thixotropic System." He then joined the Korean

Army for six months as a second lieutenant. Before he came to the United States, he

worked as a researcher for Cheil Synthetic, Inc. where he was participated in

developing a photographic film technique. He joined LSU in the fall o f 1989 and

received his Ph. D. in macromolecular chemistry in December o f 1994. His major

professor was Dr. Paul S. Russo.

185

DOCTORAL EXAMINATION AND DISSERTATION REPORT

Candidate: Daewon Sohn

Major Field: Chemistry

Title of Dissertation: Dynamic Behavior of Magnetic Latex Particles andP o lye lectro ly tes

Approved:

Major Professor and Chairman

/ Dean o: le Graduate School

EXAMINING COMMITTEE:

t-u__

Date of Examination:

1 0/27/94