175
CHARACTERIZATION OF MAGNETIC NANOPARTICLE MOBILITY IN BIOLOGICAL MATRICES AS A FUNCTION OF MAGNETIC RELAXATION MEASUREMENTS By ANA CAROLINA BOHORQUEZ A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2017

© 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

CHARACTERIZATION OF MAGNETIC NANOPARTICLE MOBILITY IN BIOLOGICAL MATRICES AS A FUNCTION OF MAGNETIC RELAXATION MEASUREMENTS

By

ANA CAROLINA BOHORQUEZ

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2017

Page 2: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

© 2017 Ana Carolina Bohorquez

Page 3: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

To my family and my beloved Venezuela No matter the distance, time or circumstance that separated us…

I love you more than anything.

Page 4: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

4

ACKNOWLEDGMENTS

I would like to thank my advisor Dr. Carlos Rinaldi for his direction and support in

completing my master degree at Universidad de Puerto Rico, Recinto Universitario de

Mayagüez and my doctoral studies at the University of Florida. This long journey has

taught me valuable lessons in both my professional and personal life. His expertise,

knowledge and excellence motivated me to learn more, and design, perform and

analyze experiments critically.

I would like to acknowledge my committee members, Dr. Brandi Ormerod, Dr.

Jennifer Andrew and Dr. Jon Dobson for their suggestions and constructive criticism of

my work. I am glad to have had the opportunity to train in confocal microscopy under

the supervision of Dr. Brandy Ormerod and her former student Aditya Asokan. I would

like to thank Dr. Andrew and Dr. Dobson for sharing their laboratory facilities.

I appreciate the opportunity given by the J. Crayton Pruitt Family Department of

Biomedical Engineering and the University of Florida for pursuing my doctoral studies in

the United States.

I am grateful for all the people that I have met since I left Venezuela. I am

fortunate to have had met many friends, mentors, mentees and colleagues that have

helped me grow as a person and as a scientist. I would particularly like to thank per

country the following people: i) In Venezuela: my mentors Dr. Rafael Belmonte and

Humberto Kum for all their guidance during the previous years. It was a bitter pill to

swallow that both of you left this world after so many years of passionate teaching while

I was abroad. I would like to thank my team “5to B” Gleidy Mendoza, Daiany Palacios,

Carlos Luis Quijada and Jhon Pulido, for supporting me in all my life transitions and I

hope our friendship evolves to a next level to plan more international meetings together.

Page 5: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

5

I would like to thank Yulyanna Carrasco for her friendship and support during my

graduate studies. ii) In Puerto Rico: I am indebted to Dr. Antonio Esteves for recruiting

me in El Salvador to pursue graduate studies at the UPRM. This decision to move to PR

changed my entire life. I thank Edwin Colon and Olga who were my hosts in PR, for

their kindness and for sharing with me all the beautiful Puerto Rican traditions. I would

like to extend my appreciation to all the lab members in The Rinaldi Lab back in PR,

specially Dr. Magda Latorre-Esteves, Dr. Mar Creixell, Dr. Adriana Herrera, Vanessa

Ayala, Dr. Jeremiah Hubbard, Dr. Lenibel Santiago-Rodriguez, Dr. Roberto Olayo and

Dr. Maribella Domench for their help and mentorship. I am very grateful to Dr. Madeline

Torres-Lugo for her technical advice and for sharing her expertise over the last few

years while working in cancer nanotechnology related projects. I would also like to thank

Dr. Karem Court and Fernando Merida for being part of my life and my scientific circle

while trying to deal together with such complicated but very challenging and rewarding

projects. Finally, iii) In Gainesville: I would like to thank Diana Dampier for helping out in

dealing with the administrative stuff at the BME department in order to get access to

essential resources during my transition from UPRM to UF. I would like to thank the

Rinaldi Lab members at UF, especially former member Tapomoy Bhattacharjee for

helping with the induction heating instrumentation and sharing insights designing

potential coil applicators for MagMED in vitro and in vivo; the undergraduate students

Francisco Ferrer, Ulises Bautista, Cory French and Jack Ke for their time and efforts in

the design and optimization of several protocols to validate MagMED in vitro and in vivo.

I also want to thank Donald Beljeri and Lunyun Su for their contribution running and

analyzing DMS measurements, and Chuncheng Yang for running and analyzing bulk

Page 6: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

6

rheological measurements for BSA solutions. I would like to extend my thanks to

Andreina Chui Lam for her assistance in the animal work, Melissa Cruz and Lorena

Maldonado for providing cobalt ferrite magnetic nanoparticles coated with PEI for the

mobility experiments, Mytheryi Unni for her contribution in trying to figure out the best

imaging conditions to perform TEM for tumor tissue samples with magnetic

nanoparticles along with helping with the magnetic characterization of the samples used

to study particle mobility. I would like to acknowledge and thank Rohan Dhavalikar for

his help, support and invaluable friendship during my graduate studies.

I also want to thank Dr. Ragy Raghed from Malvern Instruments, Jiajun Lei and

Sayali Belsare for their help in standardizing, performing and analyzing NTA

measurements using the synthesized magnetic nanoparticles. I would like to

acknowledge the efforts of Dr. Jill Verlander, Dr. Sharon Matthews and Dr. Karem

Kelley during my training in the TEM, and technical discussions about the sample

preparation and imaging during the mobility studies. I would like to thank to Dr. Dietmar

Siemann, Sharon Lepler, Chris Pampo and Dr. Lori Rice for providing guidance during

the studies using the triple negative breast cancer model to test MagMED.

Lastly, I would like to thank the entire BME family, especially Dr. Allen, Dr.

Dobson, Dr. Stabler, Dr. Judge, Dr. McFetridge, Dr. Sharma, Dr. Schmidt and Dr.

Hudalla lab members for being great colleagues and for their help in moving forward in

my projects in several situations. I would like to specially thank Brittany Jacobs, Dr.

Adam Molsalve, Maria Coronel, Irayme Labrada, Dr. Kerim Gattas and Antonietta

Restuccia. Finally, I would like to thank Dr. Paul Carney, Dr. Francisco Delgado and

Gowri Natarajan for their collaboration to test MagMED in a brain cancer animal model.

Page 7: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

7

TABLE OF CONTENTS page

ACKNOWLEDGMENTS .................................................................................................. 4

LIST OF TABLES .......................................................................................................... 10

LIST OF FIGURES ........................................................................................................ 11

ABSTRACT ................................................................................................................... 14

CHAPTER

1 BACKGROUND AND SIGNIFICANCE ................................................................... 16

1.1 Characterization of Magnetic Nanoparticle Mobility in Biological Matrices to Enhance Particle Design for Biomedical Applications ...................................... 16

1.2 The Importance of Studying Magnetic Nanoparticle Mobility in Crowded Biological Environments ................................................................................... 18

1.3 Translational and Rotational Mobility of Nanoparticles in Biological Matrices .. 19

1.3.1 Diffusion in Biological Systems ................................................................ 20 1.3.2 Frictional Coefficients for Spheres ........................................................... 22

1.3.2.1 Translational friction factor ( tf ) ...................................................... 23

1.3.2.2 Rotational friction factor ( rf ) .......................................................... 24

1.3.2.3 The Stokes-Einstein and the generalized Stokes-Einstein relations applied to complex biological environments ............................. 25

1.3.3 Colloidal Stability of Magnetic Nanoparticles in Biological Environments ................................................................................................ 29

1.4 In Situ Methods to Study Nanoparticle Colloidal Stability, Nanoparticle Mobility and Adsorption of Proteins.................................................................. 32 1.4.1 Optical Methods ...................................................................................... 32

1.4.2 Differential Centrifugal Sedimentation ..................................................... 33 1.4.3 Magnetic Measurements ......................................................................... 34

1.4.3.1 Magneto-optical birefringence ........................................................ 34

1.4.3.2 Temperature-dependent AC-susceptometry .................................. 35 1.4.3.3 Dynamic magnetic susceptibility measurements ............................ 36 1.4.3.4 Magnetic particle spectroscopy ...................................................... 37

1.5 Magnetic Relaxation Measurements to Study Mobility of Magnetic Nanoparticles in Biological Matrices ................................................................ 38 1.5.1 Analysis of Dynamic Magnetic Susceptibility Spectra to Obtain

Nanoparticle Size Distributions ..................................................................... 42 1.5.2 Analysis of Dynamic Magnetic Susceptibility Spectra to Obtain

Particle Rotational Diffusivity ......................................................................... 43

1.6 Dissertation Outline and Objectives .................................................................. 44 1.6.1 Objectives ................................................................................................ 44

1.6.2 Chapters Overview .................................................................................. 45

Page 8: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

8

2 MONITORING NANOPARTICLE-PROTEIN INTERACTIONS IN SITU BY DYNAMIC SUSCEPTIBILITY MEASUREMENTS .................................................. 48

2.1 Introduction ....................................................................................................... 48

2.2 Materials and Methods ...................................................................................... 49 2.2.1 Synthesis of CMDx with Different Degrees of Substitution ...................... 50 2.2.2 CMDx Acidimetric Titration ...................................................................... 50 2.2.3 Preparation of CMDx Coated Cobalt Ferrite Nanoparticles ..................... 51 2.2.4 Characterization of Magnetic Nanoparticles ............................................ 52

2.2.5 Dynamic Magnetic Susceptibility Measurements in Protein Suspensions .................................................................................................. 53

2.3 Results and Discussion ..................................................................................... 54 2.3.1 Nanoparticle-Protein Interactions Between ICO Nanoparticles and

Negatively Charged Proteins ......................................................................... 57 2.3.2 Nanoparticle-Protein Interactions between ICO nanoparticles and

Positively Charged Proteins .......................................................................... 59 2.4 Conclusions ...................................................................................................... 61

3 ROTATIONAL DIFFUSIVITY OF MAGNETIC NANOPARTICLES IN CONCENTRATED PROTEIN SOLUTIONS ........................................................... 71

1.1 Introduction ....................................................................................................... 71

3.2 Materials and Methods ...................................................................................... 72 3.2.1 Synthesis of Magnetic Nanoparticles....................................................... 73

3.2.2 Characterization of Magnetic Nanoparticles ............................................ 75 3.2.3 Unbuffered and Buffered, Salted and Non-Salted BSA Stock Solutions

at pH = 2.7 and pH = 7.4 ............................................................................... 76 3.2.4 Bulk Viscosity Measurements and DMS Measurements ......................... 78

3.3 Results and Discussion ..................................................................................... 80

3.3.1 Protein Solutions and Potential Aggregates/Cluster Formation ............... 85 3.3.2 Dynamic Magnetic Susceptibility Measurements and Rheological

Analysis ......................................................................................................... 88 3.3.3 Rotational Diffusion and its Deviation from the Predictions Based on

the Stokes-Einstein Relation ......................................................................... 94

3.4 Conclusions ...................................................................................................... 95

4 DETERMINING THE MOBILITY OF POLYMER COATED NANOPARTICLES IN BLOOD AND TUMOR TISSUE USING MAGNETIC RELAXATION MEASUREMENTS ............................................................................................... 114

4.1 Introduction ..................................................................................................... 114 4.2 Materials and Methods .................................................................................... 114

4.2.1 Animals and Tumor Model ..................................................................... 115 4.2.2 Synthesis of Cobalt Ferrite Magnetic Nanoparticles .............................. 116

4.2.2.1 PEI coated cobalt ferrite MNPs .................................................... 116 4.2.2.2 CMDx coated cobalt ferrite MNPs and PEG coated cobalt ferrite

MNPs .................................................................................................... 117

Page 9: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

9

4.2.3 Iron Oxide Magnetic Nanoparticles ....................................................... 117 4.2.4 Surface Modification, Amine Functionalities and Amine Quantification . 118

4.2.4.1 PEG coated MNPs – amine functionalities ................................... 118

4.2.4.2 CMDx coated MNPs and capped CMDx coated MNPs – amine functionalities ........................................................................................ 118

4.2.4.3 Quantification of amine groups in the surface of the magnetic nanoparticles ........................................................................................ 119

4.2.5 Characterization of Magnetic Nanoparticles .......................................... 120

4.2.6 Blood Collection and Tumor Harvesting ................................................ 122 4.2.7 Assessment of Particle Mobility Using Dynamic Magnetic

Susceptibility Measurements ....................................................................... 123 4.2.8 Tissue Preparation for Electron Microscopy .......................................... 124

4.3 Results and discussion ................................................................................... 125

4.3.1 Particle Mobility in PBS, Whole Blood and Tumor Tissue Using Magnetic Relaxation Measurements ........................................................... 130

4.3.2 Visualization of Nanoparticle-Tumor Interactions Using Transmission Electron Microscopy .................................................................................... 134

4.4 Conclusions .................................................................................................... 135

5 SUMMARY ........................................................................................................... 153

5.1 Estimating rotational diffusivity of magnetic nanoparticles in concentrated protein solutions ............................................................................................. 153

5.2 Determining the mobility of polymer coated magnetic nanoparticles in blood and tumor tissue ex-vivo using magnetic relaxation measurements .............. 153

5.3 Concluding Remarks ....................................................................................... 154

REFERENCES ............................................................................................................ 155

BIOGRAPHICAL SKETCH .......................................................................................... 174

Page 10: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

10

LIST OF TABLES

Table page 2-1 Hydrodynamic diameter distributions values obtained from dynamic light

scattering (lognormal volume weighted) and DMS measurements at 0.9 mg core/mL .............................................................................................................. 56

3-1 Buffer conditions employed for bulk rheology and DMS measurements of BSA at different concentrations .......................................................................... 82

4-1 Physicochemical characterization of coated magnetic nanoparticles used in this study. ......................................................................................................... 127

Page 11: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

11

LIST OF FIGURES

Figure page 2-1 Monitoring nanoparticle-protein interactions though rotational Brownian

relaxation ............................................................................................................ 63

2-2 DMS spectra of bare cobalt ferrite (CoFe2O4), showing their Debye behavior in a 60:40 %wt water:glycerol mixture. In-phase susceptibility component (○) and out-of-phase susceptibility component (●) ................................................... 64

2-3 Representative Transmission Electron Microscope (TEM) images for CMDx coated cobalt ferrite nanoparticles with different degrees of carboxylic acid substitution and lognormal core size distributions by weighted number. ............ 65

2-4 Hydrodynamic diameter distributions of the particles in the absence of proteins determined from dynamic light scattering and dynamic susceptibility measurements .................................................................................................... 66

2-5 Representative thermogravimetric analysis of CMDx coated cobalt ferrite nanoparticles. ..................................................................................................... 67

2-6 Hydrodynamic diameter distributions of magnetic nanoparticles interacting with negatively charged proteins at different protein concentrations. ................. 68

2-7 Representative dynamic magnetic susceptibility spectra and hydrodynamic diameter distributions of ICO-A particles interacting with positively charged proteins at different concentrations. .................................................................... 69

2-8 Diameter distributions of ICO-B and ICO-C particles in contact with different cationic protein concentrations at 37°C. ............................................................. 70

3-1 Particle mobility and colloidal stability in concentrated protein solutions can be studied by means of monitoring magnetic particle rotation in an alternating magnetic field and calculating the rotational diffusivity of magnetic nanoparticles ...................................................................................................... 98

3-2 Thermogravimetric analysis of oleic acid coated cobalt ferrite nanoparticles and PEG-Silane coated MNPs (n = 3) ................................................................ 99

3-3 Physical, magnetic, and colloidal characterization of the PEG-Silane coated cobalt ferrite nanoparticles. .............................................................................. 100

3-4 Stability of MNPs at pH values between 2.7 and 10 in water. .......................... 101

3-5 Particle stability at the buffer conditions used to study PEG-cobalt ferrite MNPs mobility at higher BSA concentrations ................................................... 102

Page 12: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

12

3-6 Equilibrium magnetization curves at 300 K for the PEG coated MNPs ............. 103

3-7 BSA protein solutions in acidic conditions at 4ºC. ............................................ 104

3-8 BSA solutions showing aggregates/cluster and gel formation. ......................... 105

3-9 Decreasing concentration of BSA at pH = 2.7 with an ionic strength 137 mM NaCl (sodium citrate + citric acid 0.1 M buffer) in the AC-susceptometer vials (200 µL). ........................................................................................................... 106

3-10 Normalized out-of-phase dynamic magnetic susceptibility for nanoparticles in BSA at pH = 2.7 buffered using citric acid + sodium citrate 0.1 M 137 mM NaCl ................................................................................................................. 107

3-11 Rotational diffusivity of PEG coated MNPs in acidic and basic conditions with and without the presence of salts. .................................................................... 108

3-12 Bulk rheological measurements at pH = 2.7 buffered and no salted (A-B-NS conditions) ........................................................................................................ 109

3-13 Viscosity measurements of BSA in buffered and salted conditions at pH = 2.7 and pH = 7.4 obtained via bulk rheological measurements .............................. 110

3-14 Viscosity measurements of BSA in buffered and salted conditions at pH = 2.7 and pH = 7.4 obtained via DMS measurements ............................................... 111

3-15 Comparison between viscosity values of BSA protein solutions in buffered and salted conditions at pH = 2.7 and pH = 7.4 obtained via macromolecular rheology and DMS measurements ................................................................... 112

3-16 Deviation of the Stokes-Einstein relation in BSA solutions below and above the BSA isoelectric point ................................................................................... 113

4-1 Particle mobility assessment in blood and tumor tissue ex-vivo. ...................... 137

4-2 FTIR measurements for cIO-APS, cIO-CMDx, capped cIO-CMDx and cIOPEG magnetic nanoparticles ....................................................................... 138

4-3 Representative TEM micrographs and DMS spectra for PEI, CMDx, starch and PEG coated MNPs. ................................................................................... 139

4-4 Comparison between polymer coated particles selected in this study. ............. 140

4-5 Equilibrium magnetization curves at 300 K for the PEI coated magnetic nanoparticles ................................................................................................... 141

4-6 DMS spectra and lognormal diameter distributions for thermally blocked PEI, CMDx and PEG coated MNPs in PBS, whole blood and tumor tissue ............. 142

Page 13: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

13

4-7 Initial suceptibility values normalized with respect to samples in PBS and arithmetic mean hydrodynamic diameter – PEI, CMDx, PEG coated thermally blocked cobalt ferrite MNPs .............................................................................. 143

4-8 DMS spectra and lognormal diameter distributions for thermally blocked BNF-Starch 80 nm and 100 nm MNPs in PBS, whole blood and tumor tissue . 144

4-9 Normalized initial susceptibility values with respect to PBS – BNF-Starch thermally blocked iron oxide nanoparticles ....................................................... 145

4-10 DMS spectra and lognormal diameter distributions for cIO-MNPs as-synthetized by co-precipitation method with different surface functionalities .... 146

4-11 Monitoring peak frequency displacement to the lower frequencies area upon contact with PBS, blood and tumor tissue for MNPs with magnetic relaxation mixtures ............................................................................................................ 147

4-12 TEM micrographs of PEI coated cobalt ferrite nanoparticles in tumor tissue ex-vivo. ............................................................................................................. 148

4-13 TEM micrograph of PEG coated iron oxide nanoparticles in tumor tissue ex-vivo. .................................................................................................................. 149

4-14 TEM micrographs of commercial BNF-Starch magnetic nanoparticles in tumor tissue ex-vivo. ......................................................................................... 150

4-15 TEM micrographs of CMDx coated iron oxide nanoparticles in tumor tissue ex-vivo. ............................................................................................................. 151

4-16 TEM micrographs of Capped CMDx coated iron oxide nanoparticles in tumor tissue ex-vivo. ................................................................................................... 152

Page 14: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

14

Abstract of Dissertation Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

CHARACTERIZATION OF MAGNETIC NANOPARTICLE MOBILITY IN BIOLOGICAL

MATRICES AS A FUNCTION OF MAGNETIC RELAXATION MEASUREMENTS

By

Ana Carolina Bohorquez

May 2017

Chair: Carlos Rinaldi Major: Biomedical Engineering

The success of nanotechnology-based therapies depends on the mobility and

lengths over which nanoparticles can move through biological barriers to reach their

target site. Current efforts to reduce the uncertainty in clinical translation of magnetic

nanoparticles for biomedical applications are intimately related to the study of the

absorption, distribution, metabolism and elimination patterns of these nanocarriers.

However, there is a lack of understanding of how biological barriers can reduce the

diffusion of magnetic nanoparticles due to Derjaguin, Landau, Vervey and Overbeek

(DLVO) and non-DLVO interactions, which can impact the nanoparticle’s colloidal

stability and mobility in complex milieu, increasing the risk of toxicity and compromising

intended use. This is in part because the available instrumentation to perform such

studies are based on optical approaches, which are limited to optically transparent

samples. A less explored approach to study particle mobility and stability in situ involves

the dynamic magnetization response of magnetic nanoparticles to an alternating

magnetic field, which can be used to monitor changes in the local environment of the

magnetic nanoparticles with precision and sensitivity. In this dissertation, the use of

dynamic magnetic susceptibility (DMS) measurements to study the colloidal stability and

Page 15: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

15

mobility of nanoparticles in protein suspensions, blood, and tumor explants was

established. The studies presented here are the first to assess nanoparticle mobility in

highly concentrated protein solutions, blood and tumor stroma in situ using time-

dependent magnetic susceptibility measurements. By continuously studying magnetic

nanoparticle mobility in complex milieu via DMS measurements, as a useful

assessment tool to design nanoparticles for biomedical applications without the

limitation of optical transparency in sample evaluation, innovative design features of

magnetic nanoparticles can be rationally incorporated. Considering these features one

can design new generation of magnetic nanotherapeutics with enhanced colloidally

stability and minimal interactions with blood, plasma proteins and tumor-stroma

microenvironment. We expect that this technique will contribute to assessing

nanoparticle diffusion rates, thus helping design particles with high mobility in biological

environments.

Page 16: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

16

CHAPTER 1 BACKGROUND AND SIGNIFICANCE

1.1 Characterization of Magnetic Nanoparticle Mobility in Biological Matrices to Enhance Particle Design for Biomedical Applications

It is almost a century since Heilbrunn and Seifritz [1] first used nickel

microparticles in 1920s with the aim to determine rheological properties of non-living

jellylike tissue. This was followed by Crick and Hughes [2], who in 1950 performed

uptake experiments of colloidal iron oxide particles in chicken cell embryos in order to

study intracellular rheology, revealing no toxic effects of iron oxide particles in those

cells in contrast to nickel particles. Currently, the clinical applications of iron oxide

magnetic nanoparticles (MNPs) include (i) oral and intravenous iron-dextran complex

therapy in iron deficiency anemia since 1947 [3], (ii) the use of superparamagnetic iron

oxide MNPs as a negative contrast agent for liver tumor enhancement in magnetic

resonance imaging since 1988 [4], and (iii) local magnetic hyperthermia for cancer

treatment in lymph nodes, postulated by Gilchrist et al.[5] in 1957, and currently used in

clinical trials for the treatment of glioblastoma multiforme and prostate adenocarcinoma

in Europe and United States [6]. Even today, interest in the use of iron oxide magnetic

nanoparticles continues to increase because of advances in particle synthesis,

development of suitable surface modification strategies with biocompatible polymers

and controlled bioconjugation strategies to enhance their in vivo pharmacokinetics and

biodistribution profiles of MNPs [7-9]. Also, the ability to manipulate MNPs in the

presence of an alternating magnetic field [10] along with research efforts to understand

nanoparticle-protein interactions, and the development of innovative methods for the

characterization of MNPs in biological environments [11-15] have played a key role in

engineering magnetic nanoparticles for biomedical applications. Inorganic magnetic

Page 17: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

17

nanoparticle cores embedded in several biocompatible polymers have been prepared

and characterized for potential medical applications ranging from targeted magnetic

fluid hyperthermia [16, 17], drug delivery [18], magnetically triggered cargo release [18-

20], optical and magnetic particle imaging [21], magnetic analyte capture (biomarkers,

circulating tumor cells, and human pathogens) [22-24], gene delivery [25-31],

nanowarming of cryopreserved tissue [32] and cell receptor activation [33-36].

In the biomedical applications mentioned earlier, colloidally stable, well-defined

and safe magnetic nanoparticles have been characterized in water, saline solutions,

biological buffers and complete cell culture media, but not in fluids mimicking the

complex biological environment. Therefore, there is a lack of understanding of how

MNPs interact with biological barriers such as blood and other body fluids (e.g.

cerebrospinal fluid, epithelial lining fluid or mucus, peritoneal fluid, and synovial fluid) or

with dense connective tissue structures such as tumor stroma or vitreous humor.

Understanding such interactions and the ability of MNPs to remain colloidally stable to

achieve an efficient targeting to specific tissue organs and cells vis a critical step in the

rational design of MNPs for biomedical applications. Strategies to design MNPs for use

in therapeutic applications should also be based on understanding the stability and

mobility patterns of MNPs in the biological milieu, which consists of confined and

crowded conditions of macromolecules and ions in the nanometer to micrometer range.

Principally, the stability and mobility patterns of MNPs can be altered by nanoparticle-

protein interactions and those interactions can influence immune cell uptake and

distribution in the mononuclear phagocytic system (monocytes, macrophages, and

Page 18: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

18

dendritic cells), thus affecting particle clearance and biodistribution in vivo, which directly

impacts nanotechnology-based targeted therapies.

1.2 The Importance of Studying Magnetic Nanoparticle Mobility in Crowded Biological Environments

In crowded biological environments, semi-empirical corrections considering a

self-crowding factor, , have been developed to represent experimental data and

estimate the bulk viscosity. As an example, Ross and Minton in 1977 [37] described a

generalized semi-empirical form of Mooney’s equation to quantify the viscosity of

concentrated protein solutions, shown in Equation (1-1)

0 exp

1Rheo

, (1-1)

where 0 is the viscosity of the suspending medium, is the volumetric fraction of

the suspended proteins, and estimated by Mooney to be between 1.35 and 1.91 for

hard spheres. However, this model potentially fails at high protein (> 150 mg/mL) and

salt concentrations where interactions between proteins become relevant such as

charge and hydrophobic interactions become relevant [38]. This is important to address

because [39] the shear-rate dependent viscosity in protein solutions is a complex

phenomenological phenomenon that depends on the protein molecular weight, volume

fraction, friction factor, hydration factor, pH-dependent net charge, conformational state,

and intermolecular potential between proteins.

Correlations to study the concentration-dependent viscosity of protein solutions

are valuable in the pharmaceutical industry [40-44]. To test the stability of concentrated

monoclonal antibody (mAb) therapeutics, in the order of 80 mg/mL or higher [45], which

relies on detection of protein aggregates using for example, viscosity measurements are

Page 19: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

19

made as a function of concentration, pH, ionic strength and excipient species present in

the formulation. This is in part because, the presence of protein aggregates may

severely compromise drug efficacy and safety, inducing immunogenic responses [46].

The stability of proteins at high protein concentrations under different conditions are

typically required for subcutaneous administration of mAb (<1mL - 1.5 mL) [47].

However, process control via experimental validation is time-consuming and often

requires the use of large volumes of protein suspensions that are not available during

drug discovery. For example, bulk rheology characterization requires 0.750 mL – 10 mL,

or complex instrumentation such a capillary rheometer and microfluidic devices (e.g. m-

VROC; RheoSense Inc.) require 10 µL – 750 µL of protein suspension. Thus, by

monitoring magnetic particle mobility, and demonstrating that the rotational diffusivity of

tracer MNPs in concentrated protein solutions follows the predictions of the SE relation

under specific conditions, the protein concentration-dependent viscosity can be

determined in different protein systems, and protein aggregation may be evaluated

precisely in short times using a small volume of sample (< 200 µL) via magnetic

relaxation measurements at different buffered conditions.

1.3 Translational and Rotational Mobility of Nanoparticles in Biological Matrices

Water serves as the primary milieu for life’s processes and hence, it is necessary

to understand the factors that determine rates of molecular movement in aqueous

environments [48]. Most biological processes occur in an environment that is

predominately water, with macromolecules in a crowded-like environment, surrounded

by ions. The cell, as the basic unit of life, contains 70-85% v/v water and the

extracellular space of most tissues is 99% v/v water [48].

Page 20: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

20

Nanoparticles with different arrangements, compositions and therapeutic effects

can be introduced into a biological matrix within the body in a variety of ways [49-52].

However, the success of nanotechnology-based therapies depends on the mobility and

extents to which nanoparticles can move through tissue structures to arrive at a target

site. The rates of diffusive transport of nanoparticles can vary between biological

matrices (e.g., whole blood, tissues and cells) due to macromolecular crowding effects,

which are an important but neglected aspect of the extracellular and intracellular

environment. Crowding effects cannot be neglected when blood contains about 80

mg/mL of protein or knowing that the concentration of total protein inside a cell is on

average ~250 mg/mL [53]. In such complex biological environments, mass transport

equations are fundamental for the study of nanoparticle diffusion-convection patterns.

This can be defined as follow [54]

eff

CD C C

t

ν , (1-2)

where the effect of molecular diffusion on mass transport and the contribution of

convection are linked. In non-flow systems, only the molecular diffusion accounts for the

mass conservation equation, thus, effective diffusivity effD becomes a critical transport

parameter by means of connecting the translational and rotational degrees of freedom

for a molecule.

1.3.1 Diffusion in Biological Systems

Diffusion is the random motion of the solute molecules driven by collisions with

solvent molecules because of translational and rotational kinetic energy. In a dilute

system, collisions between colloidal particles are negligible, which makes the driving

force for diffusion the thermal molecular motion of solvent molecules. Indeed, in a living

Page 21: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

21

cell diffusion is highly efficient for internal transport. Mainly, in crowded biological

environments diffusion is dominant for the transport of small molecules,

macromolecules and particles [55] such as oxygen in tissues (R = 0.12 nm), drugs in

the vitreous cavity of the eye [56], and drugs in solid tumors, for example anti-VEGF

drugs such as ranibizumab R = 2.8 nm, aflibercept R = 3.7 nm, bevacizumab R = 4.6

nm or liposomal Doxorubicin R ~ 50 nm [57].

In 1905, Einstein [58] showed that the diffusivity D of a particle in suspension in

thermal equilibrium can be related to its hydrodynamic mobility according to the with the

equation

6

Bk TD

R , (1-3)

where Bk T is the thermal energy of the system, and is the shear viscosity of the

solvent. This equation was obtained based on the assumption that the concentration

distribution of the particles is uniform. Unless an external force acts on each particle, the

diffusion flux is balanced by a negative convective flux at low Reynolds (Re) number

which, is defined as

RehydraulicD

, (1-4)

where is the density of the fluid, is the mean velocity of the fluid, and hydraulicD is the

hydraulic diameter at the section normal to the flow. Also, it was assumed that the force

is exerted along in the x-axis, and the particle concentration is dilute. In fluid dynamics,

the expression 6 R is denominated as the translational friction factor ( Tf ), which is

used to quantify the drag or resistance to particle motion in the surrounding medium. On

the other hand, when the colloidal sphere exchanges angular momentum with the

Page 22: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

22

solvent, such as when the particle probe is actively driven within the solvent by a

magnetic force in an oscillatory magnetic field, the rotational frictional coefficient ( Rf ) is

considered.

1.3.2 Frictional Coefficients for Spheres

In biological systems, convection typically transports molecules over distances

for which diffusion is too low. When a fluid is in motion the character of the flow changes

dramatically above a critical value of the Re number, depending on which the flow can

be laminar or turbulent. Fluid flow in the cardiovascular system and tissues differs from

the steady laminar flow of a Newtonian fluid mainly because blood is Non-Newtonian, its

flow is pulsatile, blood vessel entrance lengths are not long enough to form a fully

developed flow, and Re numbers range between 20-1,500. In branched vessels flow

turbulences are common at normal physiological conditions. In contrast, flow

turbulences are rare in confined biological systems over the dimension of a cell (10 µm).

Certainly, cells are typically crowded macromolecular environments in which diffusion is

highly efficient for internal transport, and is usually modeled by the Stokes equation for

small Reynolds numbers (Re << 1). In this respect, low Reynolds number flows around

a sphere are relevant in several biomedical applications, and transport within cells,

proteins, and microfluidic devices [55]. Respectively, considering a solid non-interacting

sphere of radius R surrounded by a large volume of quiescent fluid and using the

continuity equation for an incompressible flow, and the corresponding components of

the Navier-Stokes equation in spherical coordinates; we can obtain the translational

friction factor and the rotational friction factor for the sphere as follows [55, 59].

Page 23: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

23

1.3.2.1 Translational friction factor ( tf )

Assuming that the sphere slowly translates and does not rotate in steady flow,

the continuity equation is simplified to Equation (1-5)

2

2

sin( )1 10

sin

rr u

r r r

(1-5)

Neglecting variations in pressure ( p ) due to gravity around the sphere and neglecting

inertial terms in the corresponding Navier-Stokes equations, the r and components

of the Navier-Stokes equation are given by Equation (1-6) and (1-7)

2

2 2 2 2 2

1 1 2 20 sin 2 cot

sin

r r rpr

r r r r r r r r

(1-6)

2

2 2 2 2 2

1 1 1 20 sin

sin sin

rpr

r r r r r r r

(1-7)

With the boundary conditions based on the behavior of the velocity components far from

the sphere

0 ( ) cosr f r

0 ( )sing r

0

and,

0 ( ) ( ) 0r as r R with f r g r

( ) ( ) 1f r g r as r

The resulting expressions for the velocities and the pressure are

3

3

0

3 11 cos

2 2

r R R

r r

(1-8)

Page 24: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

24

3

3

0

3 11 sin

4 4

R R

r r

(1-9)

23

cos2

o Rp p gx

R r

(1-10)

At the infinity, the pressure in the uniform flow is

03cos cos

2r Rp p gR

R

(1-11)

Then, the relevant component of the shear stress is

03sin

2r r R R

(1-12)

The total force F on the sphere is the integral of stress over the whole sphere surface

2 2

2 2

0 0 0 0

cos sin sin sinz rr R r RF p R d d R d d

(1-13)

06zF R (1-14)

Finally, the translational friction factor of the sphere in a pure viscous fluid is given by

the Equation (1-15)

6Tf R (1-15)

1.3.2.2 Rotational friction factor ( rf )

Assuming that the sphere slowly rotates at a constant angular velocity around

the z-axis, the torque ZT required to maintain the sphere’s rotation defines the rotational

friction factor as z RT f . Under these conditions the Stokes equation becomes:

2

2 2 2 2

1 10 sin

sin sinr

r r r r r

(1-16)

With the boundary conditions

0

sin

as r

R as r R

Page 25: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

25

To finally obtain the velocity profile relation by a rotating sphere

2

, sinR

r Rr

(1-17)

Then, the relevant component of the shear stress is

r rr r

(1-18)

and the torque needed to integrate the tangential force ( )r r R dS exerted on the fluid

by a solid surface element dS will correspond to

sinrT R dS (1-19)

2

2

0 0

3 sin sin sinT R R d d

(1-20)

2

3 3 3

0

6 sin 8T R d R

(1-21)

To close, the rotational friction factor of the sphere in a pure viscous fluid is given by

Equation (1-22)

38Rf R (1-22)

1.3.2.3 The Stokes-Einstein and the generalized Stokes-Einstein relations applied to complex biological environments

Rotational and translational diffusion rates of small compounds, proteins and lipid

molecules undergoing Brownian motion are responsible for the completion of a wide

variety of biological functions and cellular processes. For example, it has been

demonstrated that diffusive oxygen transport is determined only by translational

diffusion in comparison with the diffusive hydrogen proton transport which appears only

be based on rotational diffusion [60]. Bigger molecules such as proteins, drugs, viruses,

bacteria and nanoparticles with a radius R in confined environments diffuse following

Page 26: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

26

the Stokes-Einstein (SE) predictions, in which the diffusion coefficient D is inversely

correlated with the viscosity of the medium and the nanoparticle’s hydrodynamic

diameter. The diffusion coefficients by translation ( T SED ) and rotation ( R SED ) are

related to the particle’s mobility as translational and rotational displacements according

to the follow equations [61]

6

BT SE

k TD

R (1-23)

38

BR SE

k TD

R (1-24)

For spherical particles moving in a medium of proportionally small molecules the SE

equations still hold. Moreover, these relations have been employed for estimating the

size of particles and biomolecules in dilute suspensions and the viscous properties of

complex fluids by a variety of characterization techniques based on, for example,

dynamic light scattering (translation) or fluorescence depolarization (rotation).

In concentrated and dilute dispersions with significant particle interactions,

particle diffusion can be slowed down due to electrostatic and hydrodynamic particle

interactions with highly structured gel-like connective tissue or crowded intracellular

environments. As a consequence, the SE relation may no longer represent the particle

diffusivity. However, the SE relation still holds for a wide range of solvents,

temperatures, and pressures with R being of the order of the molecular radius with

some exceptions at the nanoscale, for example, supercooled water [62], polymer melts

[63] and polymer solutions [64]. Indeed, Tuteja and Mackay [15] studied the diffusion

coefficients of 5 nm diameter nanoparticles in polymer melts by using X-ray photon

correlations spectroscopy, and demonstrated that the SE relation between viscosity and

Page 27: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

27

diffusion coefficient was not valid for such small particles. On the other hand, for particle

suspensions in semidilute polymer solutions, nanoparticles can be trapped by the

surrounding polymer mesh and deviate from predictions based on bulk solution

viscosities. According to Poling-Skutvik et al.[64], this is in part because of the

heterogeneities present in polymer solutions over length scales analogous to the

polymer radius of gyration ( gR ) and the correlation length ( ).

Due to the importance of understanding how particle mobility in biological

complex environments towards improving the design of nanomaterials intended for

biomedical applications such as targeted delivery of drugs in biological tissues, many

research efforts have been directed to evaluate a generalized Stokes-Einstein (GSE)

relation. GSE relations have been used to study the passive or driven diffusion by

external applied magnetic fields or optical tweezers, to probe mechanical properties at

nanoscale and microscale by correlation of a diffusional property with rheological

properties such as complex viscosity and the complex shear modulus [61]. In order to

avoid positive or negative deviations from the SE relation, the length scale of the

nanoparticle probes intended to measure the mechanical properties of the surrounding

fluid should be considerable larger than the length scale of the molecules in the fluid

[65]. Mainly, this is to facilitate that the particle probes measure the average of the bulk

rheological characteristics of the complex fluid, and can detect microscale

heterogeneities and fluctuations.

Magnetic nanoparticles (MNPs), due to their ability to be manipulated by external

applied magnetic fields, are promising to assess mechanical properties of complex

environments or detect changes in their dispersion state upon protein adsorption or

Page 28: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

28

nanoparticle-protein interactions in gel-like biological environments. This can be made

possible by monitoring the magnetic relaxation of a liquid particle system as it

approaches an equilibrium state condition after removal of the applied magnetic field.

Under specific conditions, the relaxation time can be related to the hydrodynamic

diameter of the particle probe and the viscosity of the carrier fluid and thus their

rotational diffusivity. Understanding factors affecting the diffusion of magnetic

nanoparticles in complex biological fluids using magnetic relaxation measurements is

particularly important for biomedical applications because this technique requires small

sample sizes (less than 1 mL) and can be used for optically opaque samples. The

opacity of biological complex fluids limits colloidal stability assessments of MNPs using

traditional optical techniques. Consequently, there is a need to address this, to improve

and clinically translate the use of MNPs in biomedical applications.

Several therapeutic applications of magnetic nanocarriers can be improved by

using magnetic nanomaterials with enhanced colloidal stability features to diffuse

through viscoelastic biological fluids, gels, and membranes. This can be made possible

by monitoring colloidal stability of MNP probes through magnetic relaxation

measurements in complex biological matrices at time windows comparable which those

intended for use in the following applications i) systemic delivery of targeted drug-

loaded or gene-loaded nanocarriers to treat solid tumors [66], ii) intravenous injection or

inhalation of nanocarriers for antibiotics to treat cystic fibrosis [26], iii) intraocular

injections of gene-loaded nanomedicine for the treatment of degenerative diseases in

the eye [67], and iv) delivering therapeutic agents across the intact tympanic membrane

(ear drum) to treat short term inflammatory ear conditions [68]. The complex biological

Page 29: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

29

matrices to take into account in the application scenarios described above are the blood

which is composed of proteins, electrolytes and other nano-sized molecules; tumor

stroma constituted by extracellular matrix components such as collagen, fibronectin,

hyaluronan, fibrin, and proteoglycans produced by fibroblasts in the case of solid tumors

[66]; thick mucus, rich with heavily glycosylated mucins in the airways of the lungs; and

collagen-polymer chains in humor vitreous gel from the eye [56], along with the

tympanic membrane.

1.3.3 Colloidal Stability of Magnetic Nanoparticles in Biological Environments

In contact with biological environments, nanoparticles can lose their designed

features due to electrostatic, steric and hydrodynamic interactions with macromolecules

and ions, which can lead to aggregation or sedimentation. Therefore, controlled surface

chemistry with biocompatible polymers for stabilization, and suitable bioconjugation

strategies play a fundamental role in particle colloidal stability for biomedical

applications. Indeed, the stability of a colloidal dispersion of nanoparticles is defined by

the particles remaining suspended in a liquid medium after a given experimental

observation time. A stable suspension can remain dispersed in solution for long periods

of time (days to months) without a significant change in hydrodynamic diameter and

without visible aggregation. Fundamental research using the Derjaguin-Landau-Verwey-

Overbeek (DLVO) theory to predict the total energy potentials has been employed to

provide a better understanding of the effects of polymer coating thickness on the

stability of iron oxide magnetic nanoparticles in water under different ionic strength

conditions (varying NaCl concentration) and pH, considering van der Waals,

electrostatic, steric, and magnetic interparticle forces [69, 70].

Page 30: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

30

Overall, biological environments are more complex systems than water, having

different ionic strength and pH. For this reason, extended research in colloidal stability

and nanoparticle-protein interactions after nanoparticle incubation with biological buffer

solutions and biological entities have been widely performed in the last decade [71]. For

iron oxide MNPs, colloidal stability studies have been performed in conditions that

mimic biological environments, for example, saline solution, isotonic buffers (PBS or

Hank’s balanced salt solution), incomplete and complete cell culture media, diluted

blood, serum and plasma, dilute fetal bovine serum (FBS) or fetal calf serum, diluted

cationic (histone, lysozyme) or anionic (e.g. human or bovine serum albumin, fibrinogen,

transferrin and immunoglobulins) protein suspensions, sodium bicarbonate/ MES or

HEPES buffers or even sodium citrate buffer (mimicking acidic lysosomal

compartments)[71-84]. These studies have shown protein-corona formation,

nanoparticle aggregation, nanoparticle dissolution, or irreversible desorption of

anchoring ligands [7]. Moreover, the competitive adsorption of proteins onto metal

oxide, silicon oxide, and polymer surfaces in plasma was described extensively by

Vroman [85, 86] and is therefore commonly referred to as the “Vroman effect” [87, 88].

Although the first biological barrier that MNPs will encounter after intravenous

injection is blood, there are limited studies in blood and crowded macromolecular

environments to assess colloidal stability and nanoparticle-protein interaction in situ.

Plasma proteins can absorb onto the surface of magnetic nanoparticles as a function of

particle physicochemical properties, thus causing aggregation and activating immune

response in the blood stream, and in tissue by local phagocytes. These nanoparticle-

Page 31: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

31

blood interactions may result in altered pharmacokinetics and biodistribution,

compromising magnetic nanoparticle therapeutic effects [89, 90].

According to Vogler [91], protein adsorption onto nanoparticles can be studied

using two general experimental approaches. One approach measures the fraction of

protein that remains adsorbed on the nanoparticles after an adsorbent-rinsing protocol.

Because of the rinsing process, it is expected that weakly-bound proteins are removed

and what remains are strongly bound proteins, hence representing the “hard protein

corona”. Most often nanoparticles are incubated with proteins in isotonic conditions and

are subsequently separated in order to determine which proteins, if any, were adsorbed

[92]. The most common method for isolating unbound proteins from protein-nanoparticle

complexes is centrifugation [93, 94]. Depending on the size and colloidal stability of the

nanoparticles, centrifugation times can be long and/or high centrifugal forces may be

applied, potentially leading to co-settling of un-adsorbed proteins and nanoparticles.

Because the fraction of proteins adsorbed unto the nanoparticles is determined by

spectrophotometric measurement of the protein in solution, co-settling of un-adsorbed

proteins can lead to an overestimation of the extent of protein adsorption. In some

cases this process can be further complicated due to interference by the nanoparticles

with the spectrophotometric technique being used to quantify free proteins. This is

particularly the case with magnetic iron oxide nanoparticles, where it can be alleviated

by iron quantification of the supernatant and accounting for its effect on the

spectrophotometric reading [95]. The second approach described by Vogler[91],

corresponds to characterizing in situ the proteins that are tightly and loosely bound to

the nanoparticles, the so-called hard and soft protein coronas [96].

Page 32: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

32

1.4 In Situ Methods to Study Nanoparticle Colloidal Stability, Nanoparticle Mobility and Adsorption of Proteins

At present, techniques used to perform colloidal stability and nanoparticle-protein

interaction assessments in situ of magnetic nanoparticles are based on optical methods,

sedimentation through a density gradient, and magnetic measurements. All these

methods are connected by using the Stokes-Einstein relation.

1.4.1 Optical Methods

Dynamic light scattering (DLS), photon correlation spectroscopy (PCS), zeta

potential measurements, nanoparticle tracking analysis (NTA), and fluorescent

correlation spectroscopy (FCS) are common optical methods used for colloidal stability

and interaction studies. In DLS and zeta potential measurements, the speed at which

the particles diffuse due to Brownian motion is measured by correlating the rate at

which the intensity of the scattered light fluctuates, which is proportional to the sixth

power of the particle diameter, thus making this technique very sensitive to the

presence of large aggregates. NTA uses a laser scattering microscope and a charge-

coupled device camera to record extremely dilute nanoparticle suspensions with an aim

to identify and track individual nanoparticles moving under Brownian motion and uses

the conventional mean squared displacement (MSD) method to estimate the

translational diffusivity of particles in solution. Hence the hydrodynamic diameter can be

estimated using the SE equation [97].

FCS measures the fluctuation of the fluorescence intensity produced by particles

diffusing by Brownian motion in a very small volume (femtoliter). Röcker and

collaborators [98] used human serum albumin (HSA) along with polymer coated FePt

nanoparticles decorated with the dye DY-636 for quantitative protein adsorption analysis

Page 33: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

33

and, determined that it only takes 100 seconds to form a protein corona of about 3.3 nm

of thickness. This study was carried out using 4 nM nanoparticle solutions with HSA

concentrations up to 65 mg/mL. Due to high sensitivity to large aggregates, several

autocorrelation functions were measured, averaged and analyzed to properly relate the

estimated translation diffusion coefficient with the SE equation and calculate the

hydrodynamic diameter of particles in contact with HSA. Unfortunately, drawbacks of

this technique are the use of fluorescent labels which can interfere with protein

adsorption on the nanoparticle surface [99, 100] and can be quenched in proximity of

the metal cores [101], and very low particle concentrations required for analysis [102].

For all the aforementioned optical techniques, a common limitation is the optical

transparency required for sample evaluation. This restricts the use of whole blood as

carrier fluid due to its complex nature (mixture of water, serum proteins, red blood cells

and white blood cells) and study of high particle concentrations (> 10 mMFe).

1.4.2 Differential Centrifugal Sedimentation

To study patterns of aggregation, differential centrifugal sedimentation (DCS) has

been employed to produce high resolution size distributions of particles between 10 nm

to 50 um, made of several materials whose density is different from that of the solvent

[81, 103, 104]. With knowledge of solvent viscosity and particle density, and using the

Stokes equation, hydrodynamic diameter distributions in different solvents can be

determined in situ by monitoring the time required for the particles to settle a certain

distance. The concentration of particles at each size is determined by turbidity

measurements which are converted to a weight distribution using Mie’s light scattering

theory for spheres [105]. This information is used to correlate the protein corona

thickness with the apparent size change of the nanoparticle-protein complexes

Page 34: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

34

compared to the particles without proteins. Using DCS, Amiri et al.[106] demonstrated

that protein corona thicknesses between 3-5 nm affect the relaxivity and magnetic

resonance imaging contrast efficiency of magnetic nanoparticles by evaluating in situ

protein corona formation of negatively and positively charged dextran coated MNPs in

the presence of FBS (10%) at a concentration of 7 mMFe (1 hour incubation at 37°C).

The major disadvantage of this technique is that nanoparticle-protein suspensions need

to be diluted in a sucrose gradient which may disrupt the protein corona interface, and

the turbidity data cannot be corrected based on particle density and fluid density.

1.4.3 Magnetic Measurements

Dynamic magnetic measurements can be used to evaluate changes in the local

environment and size of the magnetic nanoparticles with extreme precision, and can be

linked to changes in particle mobility due to particle aggregation or re-dispersion in

different colloidal systems. Currently, techniques such as magneto-optical birefringence,

temperature-dependent AC-susceptometry, dynamic magnetic susceptibility

measurements, and magnetic particle spectroscopy (MPS) have become relevant to

study colloidal stability or particle degradation as a function of changes in magnetization

or magnetic relaxation with specific reproducibility in biological environments. A detailed

explanation of these magnetic measurements is presented in the following sections.

1.4.3.1 Magneto-optical birefringence

Dynamic magneto-optical probing is achieved when in the presence of an

external magnetic field thermally blocked particles experience a torque, causing the

particle’s magnetic dipole to align along the field direction. Thus, the MNPs impart

optical birefringence to the surrounding medium. After removing the magnetic field, the

MNPs randomly disorient and the magneto-optical birefringence relaxes with a

Page 35: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

35

characteristic time related to the rotational diffusivity of particles dispersed according to

Debye’s equation and Perrin’s equation, respectively [107-109]. In the past, this

relaxation technique has been used to quantify antibody adsorption and particle

aggregation as a function of pH [110, 111]. In addition, Lartigue et al. [77] monitored

changes in hydrodynamic sizes of negatively charged maghemite nanoparticles (10

mMFe) incubated with rat plasma (0-100%), BSA suspensions (up to 65 mg/mL) and

Roswell Park Memorial Institute (RPMI) medium, showing that protein corona formation

as a function of time affects macrophage uptake. Although highly concentrated protein

suspensions can be assessed using this technique, the fluid carrier should be

transparent because the decay of the birefringence is detected as decreasing light

intensity. Thus, whole blood may be a challenge to assess. Also, for an increase in

signal, particles with elongated structures (shape anisotropy) are preferred instead of

spherical nanoparticles [112], which limits particle evaluation.

1.4.3.2 Temperature-dependent AC-susceptometry

To study particle degradation in complex environments, temperature-dependent

magnetic properties at low frequencies have been studied using a superconducting

quantum interference device (SQUID) based AC-susceptometry. This information has

been used to evaluate biodistribution and degradation of iron oxide magnetic

nanoparticles by monitoring the out-of-phase susceptibility component as a function of

temperature [79, 113, 114]. The temperature dependence of the in-phase and out-of-

phase susceptibility components during particle degradation has shown that the shift of

the peak temperature location towards lowers temperatures over time is associated with

particle size reduction in acidic environments. Conversely, in animal tissue samples it is

Page 36: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

36

difficult to distinguish between reductions in the particle size from a possible decrease in

particle-particle interactions or breaking of magnetic particle clusters.

1.4.3.3 Dynamic magnetic susceptibility measurements

The third option corresponds to the use of dynamic magnetic susceptibility

(DMS), in which the complex magnetic susceptibility of MNPs is measured as a function

of frequency. The response of MNPs, such as ferrimagnetic cobalt ferrite, thermally

blocked iron oxide, and iron oxide with combined relaxation mechanisms can deliver

precise information about the size distribution and local viscosity of a liquid suspension.

To achieve this from the susceptibility spectra, DMS measurements are fitted to the

Debye model of AC susceptibility by considering the polydispersity of the particles using

a lognormal size distribution. From this analysis, the hydrodynamic diameter

distributions of particles interacting with biofluids can be extracted to evaluate particle

mobility, and rotational diffusivity of MNPs.

A sensor based on measurements of rotational relaxation time of magnetic

nanoparticles was proposed by Connolly and St. Pierre in 2001 [115], wherein the

change in hydrodynamic diameter of the nanoparticles upon binding to a targeted

analyte would result in a change in the characteristic rotational relaxation time of the

particles, measurable through DMS measurements as a function of frequency. This

sensor principle was demonstrated to detect specific binding of prostate specific antigen

to the surface of magnetic nanoparticles[116] and to detect S-protein and biotinylated

T7 bacteriophage [117]. A similar concept has been applied to monitor aggregation of

nanoparticles in aqueous suspensions [118] and melt polymers [119], to monitor

gelation triggered by temperature [120], to obtain nanoscale viscosity measurements

[121, 122] and to demonstrate the existence of a critical molecular weight for a

Page 37: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

37

polyethylene glycol (PEG) melt, below which the Stokes-Einstein (SE) relation precisely

describes the rotational diffusivity of PEG coated MNPs and above which the SE

relation no longer applies [123]. Recently, Soukup et al.[124] used DMS measurements

to follow the in situ magnetic response of superparamagnetic and thermally blocked iron

oxide MNPs, stabilized electrostatically by citric acid coating, after their intracellular

internalization and posterior release by freeze-thaw lysis. The findings obtained by

Soukup at el. may have significant implications in the design of MNPs for magnetically

mediated energy delivery systems [125] and confirmed the potential of DMS

measurements in monitoring in situ mobility of MNPs in cells, a crowded

macromolecular environment.

1.4.3.4 Magnetic particle spectroscopy

As a fourth option, magnetic particle spectroscopy (MPS) measurements have

shown that changes in the magnetization of magnetic nanoparticles are linked to their

colloidal stability in cell culture media. The derivative of magnetization as a function of

the magnetic field has been measured under typical magnetic particle imaging (MPI)

conditions to determine aggregation patterns, and to provide insights for the enhanced

design of MNPs for MPI applications [126]. Also, changes in the magnetization of MNPs

were monitored to detect re-dispersion of MNP aggregates mediated by the proteolytic

cleavage with trypsin or MMP-2 proteases in cell culture supernatants. These MNP

aggregates were formed due to the high affinity between neutravidin-coated MNPs and

biotinylated peptides and served as an assay platform for detection of cancer-specific

proteases [127]. Although MPS represents a sensitive magnetic tool to study colloidal

stability of MNPs, and the current theoretical description of MPS/MPI spectra allows the

calculation of the particle core size distribution, for a more precise calculation of the

Page 38: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

38

complex MPS spectra better models including appropriate expressions for the

magnetic-field dependence of the Néel time constant are required [128].

1.5 Magnetic Relaxation Measurements to Study Mobility of Magnetic Nanoparticles in Biological Matrices

The magnetization equation is a phenomenological equation describing the

behavior of magnetic nanoparticles suspended in a nonmagnetic medium. The

magnetization vector M of the particles changes according to the magnetization

equation derived by Shliomis [129, 130]:

1

eqt

Mv M ω M M M (1-25)

Here the second term on the left-hand side corresponds to the convective

derivative of the magnetization for linear motion. On the right-hand side, the first term

represents changes in magnetization due to “rotation” of the medium and the second

term represents change in magnetization eqM via various “relaxation” processes, with

combined time constant [131].

When the fluid is at rest in a stationary unidirectional magnetic field, the

equilibrium magnetization ( eqM ) is given by the Langevin function. According to the

Langevin function, in the low magnetic field limit, the magnetization is approximately

linear with H and has constant magnetic susceptibility 0 . The relation between M and

H is expressed as 0eqM H . 0 is in general dependent on suspension composition

and temperature, in addition to particle and suspending fluid properties. This relation is

given explicitly for weakly interacting particles by [132]

2 3

00

18

d c

B

M d

k T

, (1-26)

Page 39: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

39

where is the volume fraction of the magnetic particles, 0 is the permeability of free

space, dM is the domain magnetization, cd is the core diameter, and Bk T is the thermal

fluctuation energy. This low magnetization limit is appropriate when the Langevin

parameter L is much less than unity

3

0 16

d cL

B B

M Hd mH

k T k T

(1-27)

here H is the magnetic field within the fluid, and m is the magnetic moment of a single

particle.

The commonly accepted relaxation processes for practical ferrofluid suspensions

are rotational Brownian motion, and Néel relaxation [132]. When the reorientation of the

magnetic moment of a particle suspended in a liquid is tied to the rotation of the particle

itself, the Brownian relaxation time of rotational diffusion is given by Equation

3 3

2

h hB

B B

D V

k T k T

, (1-28)

where is the viscosity of the suspended fluid and hD is the hydrodynamic particle

diameter and hV is the hydrodynamic volume.

The Néel theory of superparamagnetism explains that a spontaneous change of

the magnetization direction commonly occur in single-domain particles under the

influence of thermal fluctuations, thus leading to the following expression [133]

0

0

1exp expN

B B

KV E

f k T k T

, (1-29)

where K is the anisotropy constant, V is the volume of the particle, and 0f is the

frequency factor of the order pf 10-9 s.

Page 40: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

40

In consequence, the relative contribution of the Brownian and Néel relaxation

times is related as

1 1 1

B N (1-30)

thus, relaxation time of the magnetization of the suspension is obviously determine by

the shorter of the times N and B .

Since we considered MNPs suspended in a quiescent fluid, in the magnetization

equation (1-25) the convective and rotational terms are neglected. Also, MNPs are

assumed to align along the applied magnetic field 0 cosH H t (uniaxial magnetic

field). Considering all the previous assumptions the time dependent magnetization

equation can be summarized as follow

1

eq

dMM M

dt . (1-31)

Equation (1-31) is the magnetization relaxation equation for the particle

suspension described by the simplest Debye-like equation in which any deviation of M

from its equilibrium value eqM decays exponentially t

eqM M e

. Thus, the time-

dependent magnetization response is given by [134]

( ) ( )M t H (1-32)

The complex frequency-dependent magnetic susceptibility, ( ) , depends on its

real and imaginary components as follows: ( ) ' "i , where ' and " are the

real and imaginary components of the susceptibility

0

2 2'

1

, (1-33)

0

2 2"

1

. (1-34)

Page 41: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

41

Debye’s theory of dielectric relaxation of non-interacting polar molecules under

the influence of their rotational Brownian motion and a time-varying applied field of

frequency has been applied widely to ferrofluids [135, 136]. Specifically, Debye’s

theory for MNPs featuring Brownian-like behavior, also called thermally blocked MNPs,

relates the characteristic Brownian relaxation time B of the MNPs to the viscosity of the

suspension medium, and the hydrodynamic diameter of the nanoparticle [137].

Typically, the out-of-phase susceptibility component in a Debye plot for a Brownian

MNP reaches a maximum at the condition where 1peak in the Equation (1-34),

whereas the effective relaxation time can be obtained from the inverse of the frequency

corresponding to a peak in " and this allows to determine an average particle diameter

when the viscosity of the system is known.

Dynamic magnetic susceptibility (DMS) measurements can be employed to

determine frequency-dependent diffusion coefficient of a system of interacting Brownian

relaxing MNPs. Thus, the medium viscosity can be estimated by correlating the

susceptibility response of the MNPs with their rotational diffusivity when the magnetic

nanoparticle size distribution is known [123]. Also, when it is assumed that the

interactions between the particle probes and the polymer chains in the complex fluid are

relevant (e.g. electrostatic, steric, hydrophobic and hydrodynamic interactions within a

matrix of a particular mesh size), DMS data can be used to assess mobility of MNPs

and to determine the nanoparticle size distributions due to nanoparticles-protein

interactions in dilute protein suspensions [78]. This mobility assessment tool, knowing

the viscosity of the complex fluid or assuming a reference value ( 25water C = 0.001

Pa.s), can be linked to the colloidal stability of the magnetic nanoprobes. The technique

Page 42: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

42

can be applied to relatively small sample sizes (< 200 µL) and even optically complex

and opaque fluids, with a low concentration of MNPs (~0.06%-0.2% v/v) [78, 121,

138].The equations required to determine nanoparticle size distributions or particle

rotational diffusivity are presented in the sections which follow.

1.5.1 Analysis of Dynamic Magnetic Susceptibility Spectra to Obtain Nanoparticle Size Distributions

A superconducting quantum interference device (SQUID) based AC-

susceptometer and commercially available AC-susceptometers are commonly used to

perform dynamic magnetic susceptibility measurements by applying a small AC

magnetic field ( 0 cosH H t ) that may be superimposed on a DC field. This causes a

time-dependent rotation of a collection of MNPs where the magnetic dipoles undergoes

a rotation in the same direction as the applied field. The measured signal is obtained

through an induced voltage in a receive coil. Assuming that the particles follow a

lognormal volume distribution of hydrodynamic diameter ( )v hn D , " is defined as [78]

3

0,

2 2 6

0

" ;1

hpv hpv

v h

hpv

yn D dy

y

;h

hpv

Dy

D

3

,

1

2 2

hpv

hpv

B R DMS

D

k T D

, (1-35)

where

2

2

1 lnexp

2ln2 lnv h

gh g

yn D

D

, (1-36)

v hn D is the probability density of particles having a hydrodynamic diameter of hD , hpvD

is the volume-weighted median diameter of the distribution, and ln g is the geometric

deviation of the lognormal distribution. In Equation (1-35) 0,hpv and hpv are the volume

Page 43: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

43

weighted average initial susceptibility and the relaxation time, respectively. The DMS

spectra can be fitted to the Debye’s model in Equation (1-35) using the probability

density of the particles described in Equation (1-36), through nonlinear regression in

order to determine the lognormal volume diameter distribution and a geometric

deviation as the fit parameters of the particles in a fluid of known viscosity. The

arithmetic mean diameter, PD , and standard deviation, , of the resulting diameter

distributions can be calculated as given

2lnexp ln

2

g

P hgvD D

, (1-37)

2exp ln 1g (1-38)

1.5.2 Analysis of Dynamic Magnetic Susceptibility Spectra to Obtain Particle Rotational Diffusivity(1-39)

DMS measurements can be employed to determine the local diffusivity of

particles in the surrounding environment. Independent values of the hydrodynamic

diameter in a reference solvent (e.g. water or an isotonic buffer) can be used in the non-

linear regression to Equation (1-35) to obtain the viscosity experienced by the

nanoparticles as the fit parameter. Then, the viscosity and particle size distribution are

related to the particle rotational diffusivity by Equations (1-40) and (1-41).

2

3

0

9exp ln

2

B BR DMS v h h g

h hgv DMS

k T k TD n D dD

D D

, (1-40)

2

3

,

9exp ln

2

BDMS g

h R DMS

k T

D D

(1-41)

Through this analysis, the rotational diffusion and solution viscosity can be

effectively determined.

Page 44: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

44

1.6 Dissertation Outline and Objectives

In this work, it is hypothesized that MNPs can be engineered with enhanced

colloidal stability in biological matrices by characterizing nanoparticle-protein

interactions and mobility patterns of MNPs in complex biological environments in situ

monitoring magnetic relaxation measurements. These relaxation measurements can be

obtained by dynamic magnetic susceptibility spectra generated by the oscillatory

response of the magnetic dipoles in a small amplitude alternating magnetic field. The

complex susceptibility data is directly connected to the rotational diffusion coefficient of

the magnetic nanoparticles by applying the Debye model of susceptibility in conjunction

with the Stokes-Einstein relation.

1.6.1 Objectives

The purpose of this dissertation is to systematically study colloidal stability as a

function of polymer coated magnetic nanoparticles mobility in relevant biological buffer

solutions, protein suspensions, blood and tumor tissue ex-vivo using magnetic

relaxation measurements as a complementary characterization technique for the design

of MNPs for biomedical applications. The technique focuses on measurements of the

rotational relaxation time of MNPs to provide information of particle mobility; wherein, a

change in the local environment affecting the characteristic rotational relaxation time of

the particles can be detected through dynamic magnetic susceptibility measurements as

a function of frequency. The polymer coatings carboxymethyl dextran (negatively

charged), hydroxyethyl starch and polyethylene glycol (neutral charged polymers), and

branched polyethyleneimine (positively charged) evaluated in this work, represent

commonly used polymers for biomedical applications. Inorganic cores used were

thermally blocked magnetic nanoparticles (cobalt ferrite MNPs) and commercial

Page 45: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

45

thermally blocked iron oxide nanoparticles. Moreover, it was demonstrated that the

mobility assessment can be extended to systems of particles relaxing by Brownian and

Néel relaxation mechanisms which are typically obtained by chemical co-precipitation of

ferric and ferrous salts in alkaline medium. Specific aims were to:

1. Monitor nanoparticle-protein interactions in situ by dynamic susceptibility

measurements

2. Estimate rotational diffusivity of magnetic nanoparticles in concentrated

protein solutions

3. Determine the mobility of polymer coated magnetic nanoparticles in blood and

tumor tissue ex-vivo using magnetic relaxation measurements

1.6.2 Chapters Overview

In Chapter 2, the use of dynamic susceptibility measurements to study

nanoparticle-protein interactions in situ was demonstrated. Briefly, to illustrate the

technique data relating the effect of surface charge in carboxymethyl dextran coated

cobalt ferrite MNPs and their interaction with model anionic and cationic proteins is

presented. Proteins such as bovine serum albumin (BSA), immunoglobulin G (IgG),

fibrinogen (FIBR), apo-transferrin (TRANS), lysozyme (LYZ), and histone (HIS) were

studied in a protein concentration range between 0-10 mg protein/mL whereas BSA was

studied up to 45 mg protein/mL to mimic albumin concentration in blood. Through this

study, it was confirmed that nanoparticle-protein interactions were dominated by

electrostatic repulsion between like-charged particle-protein pairs and attraction

between oppositely charged protein-particle pairs.

Once the technique was established, measurements of the rotational diffusivity

were made for neutral polyethylene glycol (PEG) coated cobalt ferrite MNPs in

Page 46: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

46

concentrated BSA protein suspensions (0-200 mg protein/mL) at different pH, with the

aim of mimicking the crowded intracellular environment. These results were reported

and compared with the values estimated using the Stokes-Einstein relation along with

bulk rheological measurements in Chapter 3. Rotational diffusivity was shown to be

consistent with the Stokes-Einstein relation in water suspensions below and above the

BSA isoelectric point. However, deviations were observed in the presence of salts and

chelating agents.

In Chapter 4, we extended these studies based on DMS measurements to study

the colloidal stability and mobility of nanoparticles in whole blood and tumor tissue. The

polymer coatings evaluated were carboxymethyl dextran, hydroxyethyl starch, PEG

and branched polyethyleneimine (PEI). Inorganic cores such as cobalt ferrite and iron

oxide magnetic nanoparticles were used. It was confirmed that the mobility of particles

in blood and particle distribution in tumor tissue depends strongly on electrostatic

interactions and the nature of the coating.

Dynamic susceptibility measurements show that the dynamic response of MNPs

in an alternating magnetic field can be used to study colloidal stability as a function of

particle mobility in situ in highly concentrated protein suspensions, blood and tumor

tissue environments. This can take place without the limitation of optical transparency in

sample evaluation, and provides a useful assessment tool to design nanoparticles for

biomedical application with enhanced colloidally stability and minimal interactions with

blood, plasma proteins and tumor-stroma microenvironment. Finally, Chapter 5 provides

a research summary with conclusions and future directions to study colloidal stability

and nanoparticle-protein interactions in situ as a function of the mobility of polymer

Page 47: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

47

coated magnetic nanoparticles in relevant biological environments using magnetic

relaxation measurements.

Page 48: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

48

CHAPTER 2 MONITORING NANOPARTICLE-PROTEIN INTERACTIONS IN SITU BY DYNAMIC

SUSCEPTIBILITY MEASUREMENTS*

2.1 Introduction

Interactions between nanoparticles and opsonin proteins can play a determining

role in the biocompatibility and fate of nanomaterials used for in vivo applications.[139]

The competitive adsorption of proteins onto metal oxide, silicon oxide and polymer

surfaces in plasma was described extensively by Vroman [85, 86] and is therefore

commonly referred to as the “Vroman effect” [87, 88]. Depending on the

physicochemical properties of nanoparticles, proteins can adsorb to the nanoparticle

surface forming a so-called protein corona,[96] which can be either quite stable (the so-

called “hard protein corona”) or fragile (the so-called “soft protein corona”)[140] and

which potentially changes over time [141]. The adsorption of proteins onto the surface

of nanomaterials can result in their specific and/or non-specific uptake by cells [142-

144], can lead to their aggregation or to their stabilization[145], could screen functional

groups on the surface of the nanoparticles, such as targeting ligands, and prevent their

action, thus ultimately impacting their distribution and fate in an organism. As such, it is

of great importance to characterize and understand the interaction of nanoparticles and

proteins.

In this contribution, we report the use of dynamic magnetic susceptibility

measurements to study nanoparticle-protein interactions in situ. The technique consists

of measuring the rotational diffusivity of thermally blocked magnetic nanoparticles in

* Reprinted with permission from A. C. Bohorquez and C. Rinaldi, "In Situ Evaluation of Nanoparticle-Protein

Interactions by Dynamic Magnetic Susceptibility Measurements," Particle & Particle Systems Characterization, vol. 31, pp. 561-570, 2014.

Page 49: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

49

diluted protein solutions. To illustrate the technique, we studied the effect of

nanoparticle zeta potential in carboxymethyl dextran coated magnetic nanoparticles and

their interaction with model anionic and cationic proteins, such as albumin (BSA),

immunoglobulin G (IgG), fibrinogen (FIBR), apo-transferrin (TRANS), lysozyme (LYZ),

and histone (HIS) in a range of protein concentrations.

2.2 Materials and Methods

Iron (III) chloride hexahydrate (FeCl3·6H2O, 97%), cobalt (II) chloride

hexahydrate (CoCl2·6H2O, 98%), sodium hydroxide (≥98%) pellets, iron (III) nitrate

nonahydrate (≥98%), dimethyl sulfoxide (DMSO, ≥98%), chloroacetic acid (≥99%),

dextran from Leuconostoc mesenteroides with average molecular weight of 9,000-

11,000 Da, and N-hydroxysuccinimide (NHS) purchased from Sigma Aldrich (ACS

reagents) were used. The following reagents, tetrametylammonium hydroxide

((CH3)4NOH, 10%), acetic acid (99.8%, for analysis), 1-Ethyl-3-(3-dimethylaminopropyl)

carbodiimide hydrochloride (EDC), glycerol (99+% extra pure) were purchased from

Fisher scientific. 3-aminopropyltriethoxysilane (NH2(CH2)Si(OC2H5)3) (APS) was

purchased from TCI America.

The proteins selected in this study were obtained from Sigma Aldrich and used

as received without further purification. The purity percentages are detailed as follow:

bovine Serum Albumin (BSA, ≥98%), lysozyme from chicken egg white (LYZ, ≥90%),

IgG from bovine serum (IgG, ≥95%), fibrinogen from bovine plasma protein (FIBR, 65-

85%), apo-transferrin bovine (TRANS, ≥98%), and histone from calf thymus (HIS) which

was a natural mixture of histones H1, H2A, H2B, H3, and H4.

Page 50: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

50

Differences between histones correspond to the fraction lysine rich (H1),

fractions slightly lysine rich (H2A and H2B), and fractions arginine rich (H3, H4),

respectively.

2.2.1 Synthesis of CMDx with Different Degrees of Substitution

For the purpose of these studies, CMDx with nominally 5, 23, and 31 carboxylic

acid molecules per dextran chain were prepared following a procedure described by

Ayala et al.[146]. Modified polymers were prepared by carboxymethylation reaction of

dextran from Leuconostoc mesenteroides by reacting with monochloroacetic acid in the

presence of sodium hydroxide at 70 °C for 1 hour. The degree of carboxymethylation

was controlled by adjusting the concentration of sodium hydroxide. In general, 20 g of

dextran were dissolved in 100 mL of distilled water. Sodium hydroxide (NaOH) was

dissolved in 34 mL of distilled water to obtain a final concentration of 1M, 2M or 3M,

depending on the desired degree of substitution of -COOH groups per dextran chain.

The solutions were cooled in an ice bath to 4°C. Afterwards, NaOH was added dropwise

to the dextran solution. Then, 29 g of solid monochloroacetic acid was added and

dissolved using a small rotor. After the reaction, the mixture was neutralized to pH 7

with acetic acid. The product was precipitated with ethanol overnight. A solid ambar

paste was recovered, suspended in water (20 mL) and dialyzed until a conductivity of ≤

10 µS/cm was obtained. Finally, the CMDx was concentrated using rotary evaporation

and dried at 60°C. The dried sample was powdered and stored at 4°C.

2.2.2 CMDx Acidimetric Titration

After reaction, the CMDx was obtained as a sodium salt. To determine the

degree of substitution (DS) of the –COOH groups by acidimetric titration it is required to

convert the CMDx to the free acid form by washing it with an acid reagent [147]. In brief,

Page 51: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

51

1 g of CMDx was washed with 14 mL of a solution of anhydrous methanol and nitric

acid 70 %v/v. This reaction mixture was mechanically stirred at 150 rpm overnight at

room temperature. The acid reagent was removed by vacuum filtration and the CMDx

was finally washed several times with ethanol to remove traces of the acid reagent, and

dried in a vacuum oven at 60°C. To accomplish the acid titration, a solution at 1% w/v of

the CMDx in free acid form was prepared in a distilled water/acetone mixture (1:1) and 5

mL to 10 mL of NaOH 0.0121 N were added (n = 3 per each CMDx prepared). This

solution was titrated with hydrochloric acid 0.0121N, using phenolphthalein as indicator.

2.2.3 Preparation of CMDx Coated Cobalt Ferrite Nanoparticles

The cobalt ferrite particles synthetized via co-precipitation [148] were surface

modified with HNO3 2 M and hydrothermally treated with ferric nitrate according to

Gomes et al.[149] MNPs were then peptized using tetramethylammonium hydroxide

(TMAO) for further surface modification using condensation of APS molecules to

covalently attach CMDx. Indeed, peptized nanoparticles were functionalized with APS to

graft functional amine groups (-NH2) onto the nanoparticle surfaces following a

procedure described by Herrera et al.[150]. For this, peptized nanoparticles were

suspended in DMSO, 10 mL APS, 1.25 mL of water, and 100 μL of acetic acid. The

reaction mixture was mechanically stirred at 150 rpm for 72 hours at room temperature.

Nanoparticles were washed four times with ethanol, and dried at room temperature to

obtain a layer of black solid APS coated nanoparticles (ICO-APS). ICO-APS

nanoparticles were functionalized with CMDx via EDC/NHS chemistry in water [73]. This

reaction was carried out by dissolving 4 g of CMDx in 40 mL deionized water (pH 4.5-5),

with 100 mg of EDC and 60 mg of NHS. The CMDx solution was mixed with the ICO-

APS solution (0.4 g ICO-APS per 40 mL of deionized at the same pH) and stirred at 150

Page 52: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

52

rpm for 36 hours at room temperature. Finally, CMDx coated nanoparticles were

washed three times with ethanol (1:3) followed by centrifugation at 7500 rpm for 15 min,

and dried at 60°C in a vacuum oven. The nanoparticles were stored at 4°C for future

experiments. This procedure was followed for each of the prepared CMDx (CMDx-A,

CMDx-B, and CMDx-C).

2.2.4 Characterization of Magnetic Nanoparticles

To prepare samples for TEM, particles were suspended in water at a

concentration of 1 mg particles/mL, 3 µL of particle solution was placed in formvar

coated copper grids and dried in a vacuum oven for 30 min. A JEOL 1200EX

Transmission Electron Microscope (TEM) was used to image the particles and the

images were analyzed to estimate the size distribution of magnetic cores. The number

weighted mean diameter ( pgvD ) and geometric deviation ( ln g ) were determined by

fitting a histogram of nanoparticle diameters determined using image analysis (ImageJ,

distributed by NIH) to a log-normal size distribution [151].

A Brookhaven Instruments Zeta Plus was used to determine the hydrodynamic

diameter of the nanoparticles and their zeta potential. The hydrodynamic diameter of

ICO-A, ICO-B and ICO-C was measured at 37°C by suspending particles (0.9 mg

core/mL) in water and filtering the suspension with 0.2 μm nylon syringe filters. For DLS

measurement the data was interpreted using the lognormal distribution of an average of

three separate volume weighted measurements, providing a histogram of particle size

and their corresponding relative intensity. Zeta potential measurements were performed

at 37°C, at 0.9 mg core/mL, in a mixture water:glycerol (40%wt glycerol) .

Page 53: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

53

A Mettler Toledo STARe TGA/DSC1 Thermogravimetric Analyzer was used to

estimate the amount of CMDx attached to the magnetic cores. Samples were dried at

115 °C in order to remove adsorbed water and solvent traces. A temperature ramp from

25°C to 800°C was used, with a heating rate of 10°C/min in air. Samples were analyzed

in triplicate to obtain an average of the remnant weight.

A Quantum Design MPMS XL-7 SQUID magnetometer was used to measure the

dynamic magnetic susceptibility of the suspended nanoparticles at 37°C. In order to

identify Brownian behavior of cobalt ferrite magnetic nanoparticles synthesized by co-

precipitation method (ICO), dynamic magnetic susceptibility measurements of bare ICO

particles were realized before modification with CMDx. The solution concentration was

14 mg core/mL, estimated by thermogravimetric (TGA) measurements. The ferrofluid

solution was mixed with 80% wt. glycerol to perform DMS measurements.

2.2.5 Dynamic Magnetic Susceptibility Measurements in Protein Suspensions

As illustrated in Figure 2-1, adsorption of protein or protein induced aggregation

would lead to a change in the effective hydrodynamic diameter of the particles and

hence in their rotational diffusivity and Brownian relaxation time. To illustrate the

technique, we consider nanoparticles coated with carboxymethyl substituted dextrans

(CMDx) of varying degrees of carboxylic acid substitution (which we call CMDx-A,

CMDx-B and CMDx-C and which have approximately 3, 23 and 31 carboxylic groups

per dextran chain), resulting in particles with varying levels of negative zeta potential.

Here, it was hypothesized that nanoparticle rotational diffusion will be affected by

interaction with proteins in a manner consistent with electrostatic attraction/repulsion

depending on the charge of the protein of interest, and at concentrations of up to 45 mg

protein/mL.

Page 54: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

54

For nanoparticle-protein interaction studies the field amplitude used in the

dynamic magnetic susceptibility measurements was 2.0 Oe, with a frequency range

between 0.1 Hz to 1,300 Hz.

The nanoparticles were suspended at a concentration of 0.9 mg core/mL (0.2%

v/v) in a water:glycerol (40%wt. glycerol) mixture filtered with 0.2μm nylon syringe filters.

Proteins at different concentrations were suspended in particle solutions by using

sonication in an ice bath before experiments were carried out (1 min). It has been

shown that at high concentrations glycerol prevents protein aggregation, a fact that is

attributed to preferential protein hydration in water-glycerol mixtures [152, 153], hence

we can be sure that for our experiments the proteins should not be significantly affected

by the addition of glycerol. Experiments were carried out at 37°C mimicking biological

conditions.

2.3 Results and Discussion

Acidimetric titration showed that the number of carboxylic acid groups per chain

were 3.1 ± 0.7, 23 ± 1, and 31 ± 9, for CMDx-A, CMDx-B, and CMDx-C respectively.

The hydrodynamic diameter of bare cobalt ferrite nanoparticles as-synthetized by the

co-precipitation method was calculated through dynamic magnetic susceptibility

measurements (Figure 2-2) and the Debye model, resulting in a hydrodynamic diameter

of 49 nm (ln σ = 0.49) by fitting the in-phase susceptibility component and a

hydrodynamic diameter of 49 nm (ln σ = 0.46) according to fitting the out-of-phase

susceptibility component. MNPs cluster formation due to the co-precipitation synthesis

method and posterior stabilization via peptization was confirmed as a result of

observation of a larger hydrodynamic diameter relative to the physical diameter

obtained by TEM analysis.

Page 55: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

55

The core diameter of the nanoparticles after CMDx conjugation was determined

by TEM analysis, which demonstrates that the modified particles have a mean diameter

of 14 nm (with a geometric deviation ln σ = 0.45), 11 nm (ln σ = 0.47), and 11 nm (ln σ =

0.43), for particles coated with CMDx-A (ICO-A), CMDx-B (ICO-B), and CMDx-C (ICO-

C). TEM images and histograms are shown in Figure 2-3, and it can be noticed that

MNPs have a wide size distribution. Note that although particles are seen in clusters

there is space between them due to the grafted CMDx.

DLS measurements at 37 °C demonstrated that ICO-A, ICO-B, and ICO-C

particles have a volume weighted hydrodynamic diameter of 68 nm (lnσ= 0.48), 86 nm

(ln σ = 0.46) and 98 nm (ln σ= 0.44), respectively (Figure 2-4). No aggregation was

observed after 1 week of suspension stored at 4 °C. The hydrodynamic diameter of the

particles was also calculated using the DMS spectra of the particles suspended in 60:40

%wt water:glycerol mixtures and a nonlinear regression fit to the Debye model weighted

by the lognormal size distribution.

As can be observed in Table 2-1, the hydrodynamic diameter distributions

obtained from fitting the Debye model weighted by a lognormal size distribution to the

dynamic magnetic susceptibility spectra are in good agreement with the distribution

obtained from dynamic light scattering in the 60:40 %wt water:glycerol mixtures in the

absence of protein.

A concentration of 0.9 mgcore/mL (0.2% v/v of MNPs) was determined to be

optimal for the dynamic magnetic susceptibility experiments such that the in-phase and

out-of-phase susceptibilities were within the SQUID magnetometer detection limits with

no particle aggregation effects. This particle concentration was used for all subsequent

Page 56: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

56

experiments in protein suspensions. When the particles were suspended in 60:40 %wt

water:glycerol mixtures the resulting pH was close to 6, hence we expect that all the

proteins retain their surface charge (negative or positive depending on their isoelectric

points). The addition of glycerol is common practice in preparing protein solutions for

crystallography [153] and does not have an impact in protein conformation. Afterwards,

the protein experiments were designed in order to develop a characterization technique

to study nanoparticle-protein interactions in situ, and the use of buffers and salts was

avoided in order to avoid electrostatic interactions from ions that can screen

nanoparticle-protein interactions in the evaluated particle platform.

Table 2-1. Hydrodynamic diameter distributions values obtained from dynamic light scattering (lognormal volume weighted) and DMS measurements at 0.9 mg core/mL

Particles

DLS AC Susceptibility

Dh, [nm] Ln σ, [A.U] Dh Re( ' ),

[nm]

Ln σ Re( ' ),

[A.U]

Dh Im( " ),

[nm]

Ln σ Im( " ),

[A.U]

ICO-A 68 0.48 59 0.32 59 0.32 ICO-B 86 0.46 83 0.35 85 0.35 ICO-C 98 0.44 87 0.29 89 0.28

The zeta potentials of the particles at 37 °C were -30 ± 1 mV, -45 ± 3 mV, -54 ± 2

mV, for ICO-A, ICO-B, and ICO-C particles, respectively. Also, no aggregation or

settling was observed during zeta potential analysis and after 1 week of suspension in

the 60:40 %wt water:glycerol mixtures at 4 °C, indicating that the CMDx coated MNPs

are stable for at least 1 week at the experimental conditions used. The CMDx coating is

covalently bonded to the particles and it is expected to stay in place, rather than being

displaced by other macromolecules as can happen when the coating is adsorbed to the

particle surface [73, 154]. The fact that the particles used here have a stable coating

and are colloidally stabilized by steric repulsion is a key factor in interpreting the

experiments discussed below as it allows us to attribute the observed changes in

Page 57: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

57

relaxation time and hydrodynamic diameter to adsorption and aggregation induced by

the presence of proteins.

Thermogravimetric analysis of bare, ICO-A, ICO-B and ICO-C magnetic

nanoparticles indicates that the cobalt ferrite cores correspond to 93.7 ± 0.3 %, 6.9 ±

0.8 % , 13.4 ± 0.5 % and 19.4 ± 0.4 %, respectively, of the total mass per particle

(Figure 2-5).

During the experimental design, proteins that possess either an acidic or basic

isoelectric point were chosen to compare nanoparticle-protein interactions from the

point of view of electrostatic interactions. The bovine counterpart of human serum

albumin, Bovine Serum Albumin (BSA), immunoglobulin G (IgG), apo-transferrin

(TRANS), and fibrinogen (FIBR), were negatively charged at the experimental

conditions, while lysozyme (LYZ) and histone (HIS) were positively charged. These

model proteins were employed for the analysis of nanoparticle-protein interactions using

ICO-A, ICO-B, and ICO-C particles. For ICO-A particles a LYZ and HIS concentration

range between 0.1-10 mg of protein/mL was used. HIS concentration for ICO-B and

ICO-C was 10 mg of protein/mL. BSA concentration used for all particles was between

1-45 mg of protein/mL. For all other proteins, we used a concentration of 10 mg/mL.

2.3.1 Nanoparticle-Protein Interactions Between ICO Nanoparticles and Negatively Charged Proteins

It was hypothesized that negatively charged proteins would repel from the

negatively charged nanoparticles, hence there would be no changes in the Brownian

relaxation time or hydrodynamic diameter. This expectation was confirmed by the

hydrodynamic diameter distributions obtained from fitting the Debye model to the

dynamic magnetic susceptibility measurements, shown in Figure 2-6.

Page 58: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

58

Carboxylate groups in aspartic acid and glutamic acid confer negative surface

charge to BSA, IgG, TRANS and FIBR proteins in the experimental conditions

evaluated. According to Hermanson [155], both aspartic and glutamic acids contain

carboxylate groups that have similar ionization properties with pKa lower than 5. For this

reason at pH above 5, these groups are generally ionized to negatively charged

carboxylates resulting in a net negative surface charge. Thus electrostatic repulsion

between particles and these anionic proteins avoids adsorption and aggregation, and

the particles can freely rotate under the action of the alternating magnetic field. This

maybe can be the reason for not observing an increase in the hydrodynamic diameter

distributions at these experimental conditions (Figure 2-6a and 2-6b).

The hydrodynamic diameter distributions obtained from fitting the Debye model

to the dynamic magnetic susceptibility measurements for ICO-B particles in proteins are

presented in Figure 2-6c and 2-6d. As with the ICO-A particles the hydrodynamic

diameter distributions obtained from fitting the Debye model are in good agreement with

the distribution obtained from dynamic light scattering in the 60:40 %wt water:glycerol

mixtures in the absence of proteins, indicating that there is little adsorption of proteins or

aggregation induced by the presence of the proteins.

The hydrodynamic diameter distributions of the ICO-C particles seemed to

decrease slightly in the presence of anionic proteins as shown in Figure 2-6e and Figure

2-6f. From a size of 89 nm (ln σ= 0.28), hydrodynamic diameter and polydispersity was

reduced to 86 nm with ln σ= 0.29 in 1 mg BSA/mL, 87 nm (ln σ= 0.26) in 10 mg

BSA/mL, 82 nm (ln σ= 0.37) in 10 mg FIBR/mL, 84 nm (ln σ= 0.13) in 10 mg IgG/mL,

Page 59: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

59

and 85 nm (ln σ= 0.16) in 10 mg TRANS/mL from fitting the Debye model weighted by

a lognormal size distribution to the out-of-phase susceptibility spectra. These shifts in

hydrodynamic diameter relative to the diameter obtained by fitting the Debye model in

the 60:40 %wt water: glycerol mixture in the absence of proteins are relatively small. At

45 mg BSA/mL the hydrodynamic diameter obtained from fitting the Debye model to the

dynamic magnetic susceptibility was 96 nm (ln σ= 0.25), a value that is close to the

DLS measurement. However, at such high protein concentrations it is expected that the

viscosity experienced by the particles may increase[156], hence our use of the viscosity

of 60:40 %wt water:glycerol mixture in determining the hydrodynamic diameter from the

dynamic magnetic susceptibility measurements may introduce an error that tends to

over-estimate the size of the particles at high protein concentrations.

2.3.2 Nanoparticle-Protein Interactions between ICO nanoparticles and Positively Charged Proteins

In the case of interaction with positively charged proteins, we hypothesized an

increase in hydrodynamic diameter and perhaps aggregation induced by electrostatic

interactions with the negatively charged particles. In addition, it is expected that these

effects would be more pronounced with increasing negative zeta potential of the

nanoparticles. Interestingly, experiments with LYZ and HIS indicated that although both

are positively charged and have similar molecular weight their interaction with the

nanoparticles were different. Figure 2-7 illustrates these interactions for ICO-A and both

cationic proteins, in terms of the hydrodynamic diameter distributions. Figure 2-7a, 2-7b,

2-7d, and 2-7e show dynamic magnetic susceptibility spectra for selected

concentrations of LYZ and HIS. In the case of LYZ no changes in hydrodynamic

Page 60: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

60

diameter were observed at concentrations of up to 10 mg LYZ/mL (Figure 2-7c). In

contrast, we observed a decrease in hydrodynamic diameter for a concentration of 0.46

mg HIS/mL (Figure 2-7f). Above this concentration, the particles were observed to

precipitate completely after experiments. We suspect that the reason for the observed

decrease in hydrodynamic diameter is aggregation and precipitation of the larger

particles in the distribution. This was consistent with the observed decrease in the

magnitude of the low frequency in-phase susceptibility and with the observation of what

appears to be a high-frequency tail at the low end of the frequency range of the out-of-

phase susceptibility.

At higher concentrations of HIS (9-10 mg HIS/mL) the Brownian relaxation peak

reappeared even though the particles did not re-disperse completely. This observation

would be due to formation of a protein coating, which can now stabilize the

nanoparticles. For ICO-A it was observed: i) a minimum HIS flocculation concentration

of 0.46 mg HIS/mL, ii) flocculation and precipitation for concentrations between 7-9 mg

HIS/mL, and iii) partial stabilization for higher concentrations. The differences between

aggregation patterns of ICO-A particles in the presence of LYZ or HIS are due to

differences in protein structure, including distribution of the positive charges and

hydrophobic sites.

In contrast to ICO-A nanoparticles, we observed significant changes in the

dynamic magnetic susceptibility spectra of ICO-B and ICO-C particles interacting with

LYZ, as would be expected on the basis of their greater negative charge and the fact

that the protein is cationic. The hydrodynamic diameter distributions are shown in Figure

2-8. It is seen that for ICO-B and ICO-C nanoparticles the dynamic magnetic

Page 61: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

61

susceptibility spectra change significantly for concentrations of 7 mg LYZ/mL for ICO-B

and 9 mg LYZ/mL for ICO-C. As with ICO-A in contact with HIS, at higher

concentrations the Brownian relaxation peak re-appeared even though the particles did

not re-disperse completely.

We only tested dispersion of ICO-B and ICO-C nanoparticles in HIS at a

concentration of 10 mg HIS/mL (Figures 2-8b and 2-8d). For both particles, we

observed dynamic magnetic susceptibility spectra consistent with extensive particle

aggregation (Figures 2-8a and 2-8c) where for the greater the increase in hydrodynamic

diameter, the greater is the frequency shift in the imaginary component of the

susceptibility. The nanoparticles in these samples were observed to completely settle

out of solution after a few hours.

2.4 Conclusions

The experiments described here prove that dynamic magnetic susceptibility

measurements seem an appropriate in situ technique to study nanoparticle-protein

interactions that lead to adsorption of proteins or aggregation of particles. The observed

patterns of aggregation, flocculation, and some re-dispersion are consistent with

nanoparticle-protein interactions being dominated by electrostatic interactions. The

advantage of these measurements is the ability to assess changes in nanoparticle

hydrodynamic diameter in the presence of proteins without having to separate the

particles from the protein solutions disrupting the so-called protein-corona which

provides the biological identity to the polymer coated MNPs. The method is fast and

easy; it requires relatively small volume sample (<200 μl) and a minimal amount of

MNPs. Optical transparency is not required, so that the technique is applicable even in

optically complex and opaque fluids such as blood. A limitation of this approach is that

Page 62: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

62

in protein mixtures the dynamic magnetic susceptibility measurements can only

determine if the proteins result in a change in particle hydrodynamic diameter, but

cannot indicate which protein is responsible for such changes. For this reason, the

reported technique should be viewed as complementary to other approaches to

characterize nanoparticle-protein interactions. Also, collection of magnetic nanoparticles

tracers must possess a Brownian fraction of MNPs that can be fitted to the Debye

model of susceptibility [157].

For the experiments reported, the choice of water: glycerol mixtures as a carrier

fluid with viscosities adjusted was made for convenience, to make sure that the

Brownian relaxation peak would lie in the frequency range of our SQUID based AC-

susceptometer, which is limited to a maximum frequency of 1,300 Hz. For particles

suspended in aqueous solutions without added glycerol it is expected that the Brownian

relaxation peak would appears at much higher frequencies and if measurements are

carried out using a SQUID AC-susceptometer, there is a chance that the lack of a

Brownian peak would introduces some uncertainty in the measurements. However, AC-

susceptometers with much wider frequency range can be built [158] and are

commercially available [159]. A susceptometer that operates at higher frequencies (>

1300 Hz) should be suitable for similar experiments without the need for glycerol, albeit

with the disadvantage of lower sensitivity compared to SQUID based AC-

susceptometers. It is expected that the observations reported here in terms of

nanoparticle-protein interactions can be extrapolated to the same systems in aqueous

solutions without glycerol, isotonic buffers, blood and other biological complex fluids.

Page 63: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

63

Figure 2-1. Monitoring nanoparticle-protein interactions though rotational Brownian relaxation

Page 64: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

64

Figure 2-2. DMS spectra of bare cobalt ferrite (CoFe2O4), showing their Debye behavior in a 60:40 %wt water:glycerol mixture. In-phase susceptibility component (○) and out-of-phase susceptibility component (●)

Page 65: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

65

Figure 2-3. Representative Transmission Electron Microscope (TEM) images for CMDx coated cobalt ferrite nanoparticles with different degrees of carboxylic acid substitution and lognormal core size distributions by weighted number. Scale bar 100 nm

Page 66: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

66

Figure 2-4. Hydrodynamic diameter distributions of the particles in the absence of proteins determined from dynamic light scattering (volume weighted distributions) and dynamic susceptibility measurements

Page 67: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

67

Figure 2-5. Representative thermogravimetric analysis of CMDx coated cobalt ferrite nanoparticles. The first mass loss corresponds to physisorbed water

Page 68: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

68

Figure 2-6. Hydrodynamic diameter distributions of magnetic nanoparticles interacting with negatively charged proteins at different protein concentrations. a), b) ICO-A MNPs in presence of BSA up to 45 mgBSA/m, and FIBR, IgG and TRANS at 10 mg protein/mL. c), d) ICO-B MNPs in presence of BSA up to 45 mgBSA/m, and FIBR, IgG and TRANS at 10 mg protein/mL. e), f) ICO-C MNPs in presence of BSA up to 45 mgBSA/m, and FIBR, IgG and TRANS at 10 mg protein/mL

Page 69: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

69

Figure 2-7. Representative dynamic magnetic susceptibility spectra and hydrodynamic diameter distributions of ICO-A particles interacting with positively charged proteins at different concentrations. a), b) representative DMS spectra of ICO-A MNPs with lysozyme showing the Debye pattern at 0.1 mgLYZ/mL and 10 mgLYZ/ml, c) diameter distributions of ICO-A nanoparticles at different lysozyme concentrations, d),e) representative DMS spectra of ICO-A MNPs with histone showing the Debye pattern at 0.1 mgHIS/mL and 0.46 mgHIS/mL, f) diameter distributions of ICO-A nanoparticles at 0.1 mgHIS/mL and 0.46 mgHIS/mL

Page 70: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

70

Figure 2-8. Diameter distributions of ICO-B and ICO-C particles in contact with different cationic protein concentrations at 37°C. a) ICO-B MNPs in presence of lysozyme protein solutions from 0.1 mgLyz/mL to 10 mgLYZ/mL, b) loss of the Brownian-like behaviour of ICO-B MNPs in presence of histone at 10 mgHIS/mL, c) ICO-C MNPs in presence of lysozyme protein solutions from 0.1 mgLyz/mL to 10 mgLYZ/mL, and d) Loss of the Brownian-like behaviour of ICO-C MNPs in presence of histone at 10 mgHIS/mL

Page 71: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

71

CHAPTER 3 ROTATIONAL DIFFUSIVITY OF MAGNETIC NANOPARTICLES IN CONCENTRATED

PROTEIN SOLUTIONS

1.1 Introduction

Nanoparticles intended for biomedical applications must navigate biological

milieu that consists of a complex, crowded, and confined mixture of structures and

molecules that span length scales from the nanoscale to the microscale. The transport

properties of nanoparticles in such complex and crowded fluid environments is expected

to deviate from those found in the simple fluids in which they are typically characterized

in the laboratory (e.g., water, saline solutions, and simple cell culture media). As such,

the characterization of nanoparticle transport properties in fluids mimicking the

biological environment is of interest in the development of nanoparticles for biomedical

applications. Understanding the mobility of particles in complex environments such as

biological fluids and highly concentrated environments plays a pivotal role in optimizing

particle physico-chemical properties (e.g. size, shape, surface coating, graft density,

colloidal stability) for successful application. Important biological fluids such as the cell

cytoplasm and blood serum consist of concentrated (60-300 mg/mL) [53] solutions of

mixtures of proteins and other macromolecules.

The rotational diffusivity of nanoparticles can be estimated experimentally by

knowing the hydrodynamic diameter of the particles and the viscosity of the media, but

there is still concern in terms of evaluating small aggregates, nanoparticle-protein

interactions and protein-protein interactions that are typically ignored with conventional

rheological measurements. Magnetic particle rheology, which helps determine the

rotational diffusivity of the nanoparticles and hence mechanical properties of the particle

medium such as viscosity through magnetic susceptibility measurements in complex

Page 72: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

72

fluids [138], is ideally suited for highly concentrated protein solutions, where scattering

of light by the proteins might interfere with optical methods of measuring nanoparticle

diffusion.

In this contribution, we report measurements of the rotational diffusivity of

magnetic nanoparticles suspended in concentrated solutions of bovine serum albumin

(BSA), as a function of protein concentration and solution pH. The method used to

measure the rotational diffusivity of the nanoparticles is based on magnetic excitation of

nanoparticle rotation and magnetic measurement of the resulting response which is

analyzed using the Debye model. We compared the rotational diffusivity obtained using

the magnetic nanoparticles with the rotational diffusivity obtained using macromolecular

rheological measurements. Below we report on the physical and colloidal properties of

the nanoparticles used in these studies and the rotational diffusivities obtained for the

nanoparticles in concentrated solutions of BSA as a function of pH. We also report the

viscosities determined for the BSA solutions using the Stokes-Einstein expression for

the rotational diffusivity, rheological measurements in a double gap viscometer and the

hydrodynamic size of the nanoparticles.

3.2 Materials and Methods

Bovine serum albumin (BSA, Sigma-Aldrich, A7906) was dialyzed before

experiments. Nitric acid (HNO3, 0.1M), potassium hydroxide (KOH, 0.1M), phosphate

buffer saline modified without calcium chloride and magnesium chloride (PBS, Sigma-

Aldrich, D5652, 137 mM NaCl), iron (III) nitrate nonahydrate, (Fe(NO3)3·9H2O, ≥98%),

tetramethylammoniun hydroxide ((CH3)4NOH, 10%), oleic acid, Mono-methoxy PEG

(mPEG), isopropyl alcohol, sodium sulfate nonahydrate (Na2O • 9H2O), and acetic acid

were purchased from Sigma-Aldrich and used as received. 3-aminopropyl

Page 73: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

73

triethoxysilane (NH2(CH2)Si(OC2H5)3) was purchased from TCI America and used as

received. Sodium citrate dehydrate (Fisher BP226-500) and citric acid anhydrous

(Fisher BP339-500) were used as received. Diethyl ether anhydrous, 0.2 µm nylon

syringe filters, 0.22 µm polyethersufone (PES) syringe filters (MILLEX GP filter unit), 10

kDa MWCO centrifugal filters (Millipore Amicon Ultracell-15 UFC901024) were

purchased from Sigma-Aldrich. Slide-A-Lyzer dialysis flasks, 10 kDa MWCO, 250 mL

(Thermo Fisher Scientific, 87762).

3.2.1 Synthesis of Magnetic Nanoparticles

The cobalt ferrite particles synthetized via co-precipitation [148] were surface

modified with HNO3 2 M and hydrothermally treated with ferric nitrate according to

Gomes et al.[149]. Then, MNPs were peptized using tetramethylammonium hydroxide

(TMAO) for chemisorption of oleic acid to further perform a ligand exchange with PEG-

Silane. The adsorption reaction to obtain cobalt ferrite nanoparticles coated with oleic

acid was carried out as described by Hubbard et. al.[160]. Briefly, 7g of dried peptized

particles were suspended in deionized water (280 mL) and placed in a sonicating water

bath (90 seconds, Fisher Scientific Mechanical Ultrasonic Cleaner FS60) to allow

particles to suspend. Then, particle colloid was placed for 20 min in a high-intensity

ultrasonic processor (Sonics Vibra-Cell VCX 750) to break aggregates in solution. Oleic

acid (~2% v/v) was added to the solution which was again placed in the sonicated water

bath for 10 min for homogenization. This reaction mixture was transferred in a 500 mL

cylindrical reaction vessel and was stirred at room temperature for 15 min. Then,

temperature was increased to 80°C at a heating rate of 4°C/min. The reaction was

carried out at 80ºC for 1 hour and oleic acid was adsorbed onto the inorganic cobalt

ferrite cores. To remove free oleic acid, nanoparticles were washed once with ethanol

Page 74: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

74

(1:3 volume ratio) via centrifugation at 7500 rpm for 15 min. Magnetic separation was

used to discard the supernatant and collect the particle precipitate. Particles were

placed in a vacuum oven at 60 °C overnight to remove ethanol. The dried oleic acid

coated nanoparticles were stored in a refrigerator at 4°C until further use.

Mono-methoxy PEG (mPEG, 2 kDa) was converted to oxidized PEG (mPEG-

COOH) using Jones reagent [70] a strong oxidizing agent that converts the terminal

hydroxyl group in mPEG to carboxylic acid. For this, 50 mmol of mPEG was dissolved in

400 mL of acetone followed by the addition of 17 mL of Jones reagent. The mixture was

stirred at room temperature for 24 hours and quenched by addition of 5 mL of isopropyl

alcohol. To remove chromium salt by-products of the reaction, charcoal was added to

the mixture (10% w/w with respect to polymer mass). Afterwards, the mixture was

vacuum-filtered several times until a clear acetone solution was obtained. The obtained

solution was concentrated to a viscous liquid using a rotary-evaporator and then dried in

a vacuum oven at 60 °C overnight. Dried polymer was suspended in water (20 g per

10mL water) and dialyzed with a membrane 1000 Da MWCO to complete the

purification process. Further, this solution was concentrated to a viscous liquid using a

rotary-evaporator and then dried in a vacuum oven at 60°C overnight to obtain a white

sticky polymer. To synthesize PEG-Silane, mPEG-COOH MW 2 kDa was reacted with

3-aminopropyl triethoxysilane in a 1:1 molar ratio under nitrogen atmosphere at 180 °C

for 2 hours. A brown paste was obtained and stored in a desiccator. Ligand exchange

with PEG-Silane was performed as described by Barrera et. al.[70] using 200 mg of

oleic acid coated cobalt ferrite nanoparticles suspended in toluene (~ 6 mg PEG-

Silane/mL) in the presence of acetic acid as catalytic agent.

Page 75: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

75

PEG coated nanoparticles were precipitated using 500 mL of diethyl ether

anhydrous and magnetic decantation. Particles were re-suspended in 50 mL of acetone

and sonicated for 10 minutes followed by diethyl ether precipitation (150 mL) and

magnetic decantation three times to remove free polymer excess. Finally, particles were

dried at 60 °C for 2h in the vacuum oven to remove traces of organic solvents, and

stored at 4 °C for further experiments.

3.2.2 Characterization of Magnetic Nanoparticles

A stock solution of magnetic particles was prepared by dissolving dry powder

PEG-Silane coated nanoparticles in deionized water, and sonicating using a Fisher

Scientific Mechanical Ultrasonic Cleaner FS60 (15 min). This resulted in a stable and

homogeneous stock solution at a concentration of 11 mg core/mL based on thermal

gravimetric analysis. The nanoparticle solution was then passed through a 0.2 um nylon

filter twice and stored in the refrigerator at 4°C. For characterization, the particle stock

solution was diluted up to 0.5 mgcore/mL for dynamic light scattering and zeta potential

measurements. The hydrodynamic diameter and zeta potential of the PEG coated

cobalt ferrite nanoparticles was characterized using dynamic light scattering (DLS,

Brookhaven Instruments ZetaPlus), while the dynamic magnetic susceptibility was

measured using a susceptometer (Dynomag, Imego), in a frequency range from 10 Hz

to 160 kHz at an applied field amplitude not higher than 0.5 mT (5 G). Direct imaging of

the polymer coated nanoparticles was done using transmission electron microscopy

with an acceleration voltage of 100 kV (Hitachi H-7600 TEM).

The percentage of inorganic core of the oleic acid and PEG coated MNPs in dry

powder form was estimated using thermal gravimetric analysis (TGA). A Mettler Toledo

STARe TGA/DSC1 Thermogravimetric Analyzer was used to estimate the amount of

Page 76: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

76

oleic acid and PEG-Silane attached to the magnetic cores. Samples were dried at 115

°C in order to remove adsorbed water and solvent traces. A temperature ramp from

25°C to 800°C was used, with a heating rate of 10°C/min in air. Samples were analyzed

in triplicate to obtain an average of the remnant weight.

The magnetic core size and saturation magnetization of MNPs stock in water

was determined by measuring the response of the equilibrium magnetization under the

application of a DC field at 300 K in a magnetic range from 6 to -6T using a Quantum

Design MPMS-3 Superconducting Quantum Interference Device (SQUID)

magnetometer.

3.2.3 Unbuffered and Buffered, Salted and Non-Salted BSA Stock Solutions at pH = 2.7 and pH = 7.4

Proteins are considered polyampholytes, due to presence of both acidic and

basic functional groups, with pH, salt, and concentration-dependent conformations

[161]. Hence, protein dynamics are governed by the surrounding buffer pH, ionic

strength, and the volume fraction of the proteins at the studied conditions. Indeed,

protein solutions can contain protein monomers in dynamic equilibrium with reversible

clusters, and also form nonequilibrium irreversible clusters [162]. Based on this

information, the preparation of BSA protein suspensions was carried out with extreme

care using a standardized protocol based on literature review of work using protein

colloidal systems at high concentrations [46, 156, 162-166], including globular proteins

and molecular antibodies to find the optimal approach. Also, observable physical

changes were documented to connect these experimental annotations with further

results using bulk rheology and DMS measurements.

Page 77: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

77

Solutions were prepared between 4°C-10ºC during most of the preparation

stages. Sterile filtration of buffers and protein stock solutions, and dilutions were carried

out in a laminar flow hood. Also, according to the supplier, the solubility of BSA in water

is 40 mgBSA/mL, thus BSA was dissolved under the buffered conditions tested in a

concentration of ~35 mgBSA/mL which is lesser than the solubility value. The pH values

selected in this study for DMS and bulk rheology measurements were below and above

the BSA isoelectric point (~ pI = 5.1)[46, 161]. A three-step procedure was followed to

prepare the concentrated BSA stock solutions at different buffer conditions.

First, a BSA solution at 35 mgBSA/mL (~480 mL) was prepared by dissolving

protein powder at different buffered solutions that were pre-filtered with 0.22 um bottle

top filters (PES membrane). Complete dissolution of the BSA powder in appropriate

buffer conditions occurred with occasional and gentle stirring at 4ºC for about 1 day

(quasi-quiescently). Briefly, ~17 g of BSA (added to the buffer slowly during about 1

day) were suspended in i) ultrapure deionized water at adjusted pH of 2.7 and 7.4,

respectively using KOH 0.1 M and HNO3 0.1 M (unbuffered and non-salted), ii) ultrapure

deionized water with 137 mM NaCl pH = 2.7 (unbuffered and salted), iii) PBS pH = 7.4

(buffered and salted), iv) sodium acetate + citric acid 0.1 M pH=2.7 (unbuffered and

non-salted), and v) sodium acetate + citric acid 0.1 M + 137 mM NaCl pH=2.7

(unbuffered and salted).

Second, protein suspensions per buffer condition were divided into two dialysis

flasks (10 kDa MWCO), and exhaustively dialyzed into each buffer condition to reduce

any ionic impurities or remove fatty acid impurities left in the BSA overnight at 4ºC. The

Page 78: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

78

pH was monitored and adjusted using KOH 0.1 M and HNO3 0.1 M, if it was required

during the dialysis process.

Third, the BSA solutions were subsequently filtered through 0.22 um syringe

filters (PES membrane) to remove any undissolved material, and concentrated by

ultrafiltration with Amicon devices (100 kDa MWCO) using a fixed angle rotor centrifuge

(Eppendorf 5430R – Rotor: F-35-6-30) at 5,000 rcf and 10˚C for about 1 day, with gentle

pipetting every 30 min to avoid protein caking in the membrane due to concentration

polarization. Then, protein solutions were centrifuged at 5,000 rcf for 30 min at 10ºC to

keep the supernatant as the final protein stock solution. This centrifugation step was

performed to remove residual particulates, and microscopic bubbles immediately before

protein concentration determination, DMS measurements and bulk viscosity

measurements. Protein concentration determination was performed with the collected

supernatant after centrifugation by using A280 method with an extinction coefficient for

BSA of 0.667 cm2/mg for a 1% dilution [46]. BSA stock solutions were diluted to the

desired concentration with the respective buffer (8 points per buffer condition), and pH

was monitored after measurements up to 24 hours. After dilution, protein concentration

were again validated by using the A280 method.

3.2.4 Bulk Viscosity Measurements and DMS Measurements

For bulk rheological measurements, 7.5 mL of the diluted protein solutions were

mixed with 2.5 mL of the appropriate buffer per condition to obtain a final target BSA

concentration. DMS measurements were performed with diluted BSA suspensions (300

µL) mixed with a freshly prepared stock solution of MNPs in the corresponding buffer

condition (100 µL). This resulted in a total particle concentration of ~0.007% v/v within

the same target BSA concentration used for rheological measurements.

Page 79: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

79

A stress-controlled rheometer MCR301 (Anton Paar) with double-gap (GP)

geometry was used to perform bulk rheological measurements. The viscosity of each

sample was measured twice at a constant set temperature of 25°C. Every time the

sample went through one cycle of shear rate sweep ramping up from 1 s-1 to 1000 s-1

and then ramping down from 1000 s-1 to 1 s-1. The average value was obtained by

calculating the average viscosity of the shear rate sweep of two independent

measurements per sample in the linear viscosity range between 20 s-1-1000 s-1.

The dynamic magnetic susceptibility of protein solutions containing MNPs was

measured using a susceptometer (Dynomag, Imego) in a frequency range from 10 Hz

to 160 kHz at an applied field amplitude not higher than 0.5 mT (5 G). Measurements at

different protein conditions were performed in triplicate at 25°C with 200 µL of protein-

nanoparticle solutions. The rotational diffusivity of the PEG coated MNPs was estimated

by assuming that the SE relation for the rotational diffusivity applies in BSA solutions,

and that the particle’s hydrodynamic diameter in the BSA solutions are the same as

those measured in the buffer solutions without proteins using DMS measurements,

which are in close agreement with DLS measurements.

As illustrated in Figure 3-1, DMS measurements represent a unique technique to

evaluate the rotational dynamics of magnetic nanoparticles in complex biological

environments. This is accomplished by sensing the particle rotation opposed by the fluid

resistance using time-dependent magnetic torques through the application of an

alternating magnetic field. Thus, the resistance due to molecular crowding would lead to

a change in the Brownian relaxation time of the nanoparticle, which is inversely

proportional to the rotational diffusivity of the nanoparticle. Protein adsorption or protein-

Page 80: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

80

induced aggregation, also can lead to a change in the Brownian relaxation time of the

particle due to a change in the effective hydrodynamic diameter.

3.3 Results and Discussion

In the experiments reported here, polyethylene glycol (PEG) was chosen as the

polymer coating grafted onto the inorganic cobalt ferrite cores, due to its relevance in

biomedical applications for preventing the adsorption of proteins via steric hindrance,

with a typically neutral charge, which reduces electrostatic interactions. Thus, PEG

coating was used to reduce adsorption of bovine serum albumin (BSA) onto the surface

of MNPs at high protein concentration and to obtain the frequency-dependent diffusion

coefficient of a system of interacting Brownian particles using dynamic magnetic

susceptibility (DMS) measurements. The physical characterization results of the PEG

coated MNPs are presented from Figure 3-2 to Figure 3-5. TGA measurements (n = 3)

presented in Figure 3-2, demonstrated that approximately 82% ± 1% of the weight loss

represents the PEG layer onto the surface of the cobalt ferrite nanoparticles.

The mean core diameter of the nanoparticles, estimated by fitting the number

weighted diameter histogram obtained using TEM to a lognormal size distribution was

14 ± 1 nm. The size distribution histogram (weighted by volume) corresponding to

Figure 3-3a is included in Figure 3-3c for comparison with the hydrodynamic size

distributions estimated from DMS and DLS measurements. Despite particle

polydispersity, well separated particle clusters of ~40 nm in diameter can be recognized.

Particle polydispersity and cluster formation is a result of synthesis by the co-

precipitation method, and after the surface modification procedure using PEG-Silane

polymers synthesized by the method described by Barrera et al. [70, 167] due to the

formation of APS oligomers.

Page 81: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

81

DMS spectra (Figure 3-3b) shows that the in-phase and out-of-phase

components of the magnetic susceptibility cross almost at the peak of the out-of-phase

susceptibility, and that both approach zero at high frequency, as expected for particles

with a single relaxation mechanism of response to magnetic fields following the Debye

model. The Brownian peak, characteristic of thermally blocked MNPs, was well defined

regardless of the polydispersity and quasi-spherical shape of the cobalt ferrite MNPs

within clusters. This is because of the well-defined polymer shell covalently bonded to

the magnetic nanoparticles after the ligand exchange procedure, which provides steric

repulsion in the particle system. In Figure 3-3c (solid black line), the arithmetic mean

diameter and the standard deviation of the hydrodynamic size distribution estimated

from fitting the DMS spectra to the Debye model was 44 ± 10 nm.

DLS measurements were used to obtain the hydrodynamic diameter distribution

of the nanoparticles and for further validation of the DMS measurements described

above. As can be seen in Figure 3-3c (dashed red line), the lognormal volume-weighted

distribution is broader than the hydrodynamic distribution obtained by fitting the DMS

spectrum to the Debye model. At a pH of 7.4, the arithmetic mean diameter and the

standard deviation of the particles was 37.0 ± 15 nm, which is lower than the

hydrodynamic diameter obtained from DMS measurements, likely due to the DMS

diameter being influenced by larger particles. The difference in the size distributions

obtained by TEM in comparison with DMS and DLS values is due to the presence of

inorganic clusters embedded in dense PEG-Silane brushes.

The characterization of the DMS spectra and hydrodynamic diameter distribution

of the magnetic nanoprobes in water over a wide range of pH showed that the particles

Page 82: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

82

are colloidally stable, relax through the Brownian magnetic relaxation mechanism, and

experience negligible changes in hydrodynamic diameter over the studied pH range

based on lognormal weighted distributions (Figure 3-4a).

Figure 3-4b shows that DLS distributions remain unchanged across pH values

between 2.7 and 10.0 in water with exception at pH = 4.7, consistent with the

observation that the DMS spectrum is unaffected by pH. Table 3-1 shows a summary of

the buffer conditions used to prepare BSA suspensions at different protein

concentrations below (pH = 2.7) and above (pH = 7.4) the isoelectric point of BSA (~pH

= 5.1). Note that each condition has a sample name that will be used consistently

throughout this entire chapter.

Table 3-1. Buffer conditions employed for bulk rheology and DMS measurements of BSA at different concentrations

Sample name

pH Solvent Buffer Ionic strength DMS – Hydrodynamic size (lognormal weighted by volume)

A-NB-NS 2.7 Water n.a 0 mM NaCl 68 nm (lnσ = 0.47) A-NB-S 2.7 Water n.a 137 mM NaCl 38 nm (lnσ = 0.27) A-B-NS 2.7 Water Citric acid

+ sodium citrate 0.1 M

0 mM NaCl 39 nm (lnσ = 0.22)

A-B-S 2.7 Water Citric acid + sodium citrate 0.1 M

137 mM NaCl 74 nm (lnσ = 0.44)

N-NB-S 7.4 Water n.a 137 mM NaCl 68 nm (lnσ = 0.33) N-B-S 7.4 Water Phosphate

buffered saline

137 mM NaCl 55 nm (lnσ = 0.36)

Sample names were denoted according to the following order: A and N for acidic (~pH = 2.7) and neutral conditions

with respect to the BSA isoelectric point (~pH = 7.4) – NB and B were designated for no buffered (unbuffered) and

buffered conditions – NS and S were intended for no salted or salted (137 mM NaCl) conditions.

Page 83: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

83

The colloidal stability analysis of nanoparticles in the buffer conditions selected

for protein suspension were performed using DLS and DMS measurements. There were

observed changes in hydrodynamic diameter distributions in selected buffers and ionic

strength conditions, as seen by a displacement of the Brownian peak by up to one order

of magnitude in Figure 3-5. However, PEG coated MNPs exhibited Brownian relaxation

even though slightly aggregated in buffered and salted conditions. Figure 3-5a and 3-5b

are representative examples of DMS spectra at pH = 7.4. Agreement between DLS and

DMS measurements in all the buffer conditions tested was observed. MNPs did not

precipitate over a period of 1 week, except for particles suspended in citric acid +

sodium citrate 0.1 M 137 mM, which precipitated after 72 h. These results confirmed

that the particles synthesized are suitable to study their rotational dynamics at different

BSA concentrations. This is because, DMS measurements are conveniently short (~20

min) and studies were carried out per buffer condition in triplicate in about a day.

Figure 3-5d and Figure 3-5e are representative examples of hydrodynamic

diameter distributions of PEG coated nanoparticles in neutral conditions, whereas in the

presence of salts at pH = 7.4 MNPs have shown a slight increase in hydrodynamic size

in contrast with deionized water at pH = 7.4 (Figure 3-4b). The close agreement

between the hydrodynamic diameters obtained by DLS and DMS measurements

indicates that the magnetic dipoles in PEG coated particles suspended at the buffer

conditions in this study, respond to an alternating magnetic field by physical particle

rotation.

Despite the presence of aggregates, PEG coated MNPs remain with a

hydrodynamic diameter less than 100 nm. The existence of this aggregation pattern can

Page 84: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

84

be explained by electrostatic interactions between free unprotected amine groups on

the surface of the PEG coated MNPs, and chloride, citrate and phosphate ions. Indeed,

the pKa of primary amines is usually high (pKa = 10), thus at pH values under pH = 10

primary amines from unreacted APS are protonated contributing to an overall positive

surface charge in the MNPs. First, zeta potential measurements at pH = 7.4 revealed

that PEG coated MNPs are slightly positive ~ 11 mV. This positive value did not change

significantly in a wide pH range (from pH 2 to pH 10), which confirms the presence of

free amine groups. Second, sodium chloride (NaCl) and PBS constituent ions may have

a screening effect on the surface charge of the MNPs because chloride and phosphate

ions in water can interact electrostatically with protonated amine groups. Finally, citrate

ions act as buffering agents, acidifiers, and chelators with metals forming coordination

compounds, which make them suitable for example, in blood anticoagulation

applications. Interestingly, citrate ions are known for being the least chaotropic ions in

the Hoffmeister series for salting or desalting proteins [168] but may favor bridging

interactions and selective particle aggregation due to electrostatic interactions of

exposed protonated amine groups onto the surface of PEG coated MNPs, and

carboxylate groups in citrate buffer used in this study. These interactions can lead to

particle hydrodynamic diameter increases. Also, citrate ions are well known as chelating

agents contributing on the degradation kinetics of MNPs which can compromise particle

stability if the polymer coating onto the magnetic nanoparticles is not stable or

covalently attached. In our experiments, particle degradation was not found in acidic

environments as the MNPs were coated with covalently grafted PEG brushes. However,

an increase of particle hydrodynamic diameter was observed in a time window of 72 h in

Page 85: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

85

acidic environments (pH = 2.7) in contrast to neutral environments (pH = 7.4), which

agrees with the idea of electrostatic interactions favor bridging interactions and selective

particle aggregation.

Equilibrium magnetization measurements at 300 K for aqueous PEG coated

MNPs, demonstrated superparamagnetic behavior (Figure 3-6) in the particle stock. By

fitting the data to the Langevin function using the procedure established by Chantrell

[169], the magnetic diameter was 14 ± 6, while the volume fraction ( ) of the dilute

stock solution was 0.02%.

For clarification, to study particle mobility and the mechanical properties of highly

concentrated protein solutions depending on the buffer and salt concentrations, the

rotational diffusivity will be estimated using the hydrodynamic size distribution obtained

via DMS measurements in each buffer condition without proteins added, which will be

used below to interpret DMS measurements in concentrated BSA solutions using the

Debye model of susceptibility along with the Stokes-Einstein relation.

3.3.1 Protein Solutions and Potential Aggregates/Cluster Formation

In this study, biophysical characterization of the BSA protein suspensions was

assessed by quantifying BSA concentration (A280) in BSA protein solutions, and

measuring the bulk viscosity in each protein solution. Following it, the particle mobility of

MNPs via particle rotational dynamics using alternating magnetic fields was studied.

Potentially, this technique will allow sensing the changes in the surroundings of the

MNPs at nanoscale due to protein aggregation, protein crowding, and/or protein

adsorption onto the surface of the magnetic nanoparticles because of DLVO or non-

DLVO interactions.

Page 86: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

86

Figure 3-7 shows protein suspensions in acidic conditions at ~35 mgBSA/mL

after dialysis, and their respective pH stability graphs about 1 day of observation.

Buffered acidic conditions in the presence of citric acid + sodium citrate 0.1 M (pH = 2.7)

showed opalescence because of protein aggregation. This physical observation agrees

with Sarangapati et. al.[161], who observed for BSA protein solutions up to 200

mgBSA/mL in citric acid + sodium citrate buffer (pH = 3) at low ionic strength (20 mM),

changes in the Small-Angle Neutron Scattering (SANS) data attributed to BSA

conformational changes.

Before centrifugal concentration in Amicon devices (100 kDa MWCO), samples

were filtered to remove protein aggregates, as shown Figure 3-8a to Figure 3-8d. During

protein concentration, it was observed that BSA gels were formed at the bottom of the

centrifugal filters. Thus, these gels were separated out from the concentrated BSA

stocks solutions (citric acid + sodium citrate 0.1 M) by centrifugation. As a

representative example, Figure 3-8e shows the supernatant collected from a BSA stock

solution buffered in citric acid + sodium citrate 0.1 M salted with 137 mM NaCl, and

Figure 3-8f shows the bottom left as a BSA gel. It was observed that without the

addition of salts the BSA gel became more solid-like and homogeneous. We therefore

expect to observe a difference in the mobility patterns of the MNPs in these BSA protein

suspensions.

During protein sample preparation, it was observed that at acidic conditions (~

pH = 2.7) without NaCl, gelation of BSA was triggered by the addition of citric acid +

sodium citrate 0.1 M buffer (A-B-NS condition). This was detected under visual

inspection during protein concentration at low temperatures using centrifugal filtration

Page 87: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

87

devices where BSA concentrations higher that 100 mg/mL were obtained. In contrast,

salted protein solutions in citric acid + sodium citrate 0.1 M buffer (A-B-S condition)

showed a reduced gel formation, which means that gelation can be inhibited by addition

of NaCl at physiological concentrations (137 mM). On the other hand, acidic

suspensions of BSA in ultrapure deionized water (A-NB-S and A-NB-NS conditions) did

not show this pattern of gel-like conformation of BSA in solution.

Esue et.al.[170] have shown carboxylate-dependent gelation of monoclonal

antibodies, in which carboxylates could act as a bridge to crosslink antibody monomers.

However, they stated that for the monoclonal antibody (mAb) studied, gel assembly was

not activated by high ionic strength but with multivalent carboxylates. These

observations were made based on sodium citrate as a multivalent carboxylate. In recent

work, Esue and collaborators [171] found that positively charged histidine play a pivotal

role in the thermodynamics of interactions between citrate and mAbs, followed by other

hydrophobic and aromatic amino acid residues that may induce assembly and

stabilization of molecular structures leading to gelation of mAbs.

The findings reported by Esue and collaborators with monoclonal antibodies

[170, 171] may be extended to globular proteins such as BSA. This is in part because,

the gelation patterns observed in our experiments agreed with their results. Based on

our physical observations, we can suggest that the citrate buffer has a determinant role

in the gelation of BSA, which can be reduced by addition of sodium chloride like

previous studies with mAbs have shown. The buffered conditions tested in our studies

lead to changes in the mechanical properties of concentrated BSA suspensions, due to

conformational changes of BSA in solution (solution pH above BSA pI) to BSA gel-like

Page 88: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

88

solid (solution pH below BSA pI). By monitoring conformational changes of BSA in

solution using dynamic magnetic susceptibility measurements, our qualitative

observations will become quantitively relevant. Hence, we will predict changes in the

rotational diffusivity of MNPs in BSA solutions and tie those variations with changes in

the mechanical properties of BSA suspensions such as BSA protein solution viscosity.

3.3.2 Dynamic Magnetic Susceptibility Measurements and Rheological Analysis

Dynamic magnetic susceptibility measurements performed for PEG coated

MNPs in BSA suspensions indicated the particles had a single magnetic relaxation

mechanism consistent with the Debye model for all the buffer conditions tested. This is

consistent with the nanoparticles remaining well dispersed and colloidally stable in all

the protein solutions. Figure 3-9 shows a representative picture of two dilution series of

concentrated BSA from 103 to 0 mg/mL without (top image) and with PEG coated

MNPs (bottom image) after 48 h in acidic conditions with an ionic strength of 137 mM

NaCl (citric acid + sodium citrate 0.1 M). By simple observation, none of the

nanoparticle-BSA solutions showed evidence of aggregation or precipitation for about

one week, with the only exception being PEG coated MNPs particles that precipitated in

the case when no protein was added (0 mgBSA/mL) at 72 h as shown Figure 3-9. In

other buffer conditions, all solutions remained brilliant and with no evidence of

remarkable particle or protein aggregation or sedimentation.

As seen in Figure 3-10, the frequency corresponding to the peak in the out-of-

phase susceptibility decreases with increasing BSA concentration (A-B-NS condition).

This is consistent with an increase in the particle relaxation time and corresponding

decrease in rotational diffusivity. However, this reduction in rotational diffusivity based

on macromolecular crowding effects can also occur because of particle aggregation or

Page 89: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

89

protein adsorption. We discarded particle aggregation because Debye-like features

remain in the DMS spectra of MNPs in all the protein conditions tested. To rule out a

protein adsorption effect which can lead to aggregation and precipitation of MNPs, the

initial susceptibility value for all the obtained DMS spectra remained almost constant.

This means that the magnetic volumetric fraction of PEG coated MNPs was the same at

different protein concentrations per buffer condition and thus particles were able to

respond by physical rotation to an externally applied magnetic field.

DMS measurements are consistent with a decrease in rotational diffusivity with

increasing concentration and decreasing pH value from 7.4 to 2.7. We can infer that the

decrease in peak frequency as concentration increased indicates that the local

environment around the magnetic nanoparticles is becoming more crowded, resulting in

hindered particle rotation. Then, if the Stokes-Einstein relation for the rotational

diffusivity is assumed to apply and the nanoparticle’s hydrodynamic diameter is

assumed to be the same in all the protein solutions, this decrease in diffusivity can be

interpreted as an increase in the viscosity of the solution in which the nanoparticles are

suspended.

However, a dramatic reduction in the rotational diffusivity of the MNPs was

observed in sodium citrate buffer without added salt (A-B-NS condition). In Figure 3-11

(filled blue squares), this severe reduction in rotational diffusivity can be linked to a

change in the BSA conformation, which produces a gel-like BSA structure. This is

because of the presence of citrate, a multivalent carboxylate ion which can trigger

crosslink and bunding of protein networks. In the DMS spectra of BSA solutions at

acidic conditions buffered with citric acid + sodium citrate buffer 0.1 M without salts (A-

Page 90: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

90

B-NS condition), particle mobility is restricted probably due to interfilamentous

interactions or steric hindrance as was shown in recent research work using spindle-

shaped MNPs in hydrogels samples [172]. It is interesting to note that this pattern is not

present at the same buffered conditions with addition of 137 mM NaCl (A-B-NS

condition). Most probably, this is because salts can screen the carboxylate ions in

solution, reducing the formation of bridges between proteins.

It is important to point out that gelation of BSA is commonly completed at high

temperatures with the presence of salts over a wide range of pH. However, in our BSA

experiments, acid-induced cold gelation of BSA between 4°C-10°C was triggered by

sodium citrate ions without the need of an increase in temperature. As an example,

acid-induced cold gelation at 25°C of globular proteins such as whey protein isolate

(WPI) has shown the formation of a protein network by physical interactions, which is

subsequently stabilized by the formation of disulfide bonds [173]. Physical interactions

can occur due to protein conformational changes – α-helices changing into β-

aggregates, and disulfide bond formation is specific for each globular protein depending

on the number of cysteine groups per protein chain. Indeed, this acid-induced cold

gelation of WPI is not triggered by the crosslinking of citrate ions, and it seems to be

relevant for comparisons with our BSA results at unbuffered acidic conditions tested (A-

NB-NS and A-NB-S conditions).

The rotational diffusivity ( ,R DMSD ) was calculated first by fitting the DMS spectra to

the Debye model weighted by the known lognormal size distribution at zero protein

concentration to obtain the viscosity experienced by the particles at the tested

Page 91: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

91

conditions, and second using the obtained viscosity value along with the lognormal size

distribution at zero protein concentration with the Stoke-Einstein equation.

Examination of the rotational diffusivity patterns in Figure 3-11 suggests that BSA

gel-like solutions in citric acid + sodium citrate 0.1 M salted (A-B-S condition) generated

restriction of MNPs mobility upon application of an alternating magnetic field for all the

acidic conditions tested (ionic screening), in comparison to BSA gel-like solutions in

citric acid + sodium citrate 0.1 M no salted (A-B-NS condition), which showed the

highest restriction in particle mobility (carboxylate bridges).

Gelation of gelatin solutions triggered by temperature has been studied by

Barrera and collaborators [120] through DMS susceptibility measurements using PEG

coated cobalt ferrite MNPs suspended in gelatin from bovine skin type B. This was

possible by monitoring sharp transitions in the out-of-phase component of the complex

susceptibility, and correlating them to observations of gelation through conventional bulk

rheology. The studies reported here demonstrated that tighter three-dimensional gels

restricted rotation and eventually trapped the MNPs. This was observed at the

nanoscale when no Debye-like peak was observed at 10°C, which was below the

gelation temperature for gelatin from bovine origin. Then, translating these observations

to our BSA studies, the gelation network formed in the BSA suspensions was not

entangled enough to totally restrict particle rotation, for this reason well defined Debye-

like peaks were observed during gelation triggered by citrate ions with and without

added salt.

The bulk rheological measurements ( Rheo ) made for BSA protein solutions were

performed at 25ºC showing apparent Newtonian behavior at shear rates between 10-

Page 92: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

92

1000 1/s. This is a characteristic pattern in the rheology of surfactant-free protein

solutions previously demonstrated by Sharma et al.[156] in BSA protein solutions.

Figure 3-12 shows representative flow curves of BSA protein solutions at acidic

conditions in citric acid + sodium citrate 0.1 M not salted.

An increase in viscosity was observed as a function of protein concentration in all

the buffer conditions using bulk rheology. Indeed, Figure 3-13 indicates that the solution

viscosity increases more rapidly at pH = 2.7 than at pH = 7.4. This could be due to

changes in the conformation of BSA at low pH values, resulting in changes in protein-

protein interactions due to cold-acidic gelation at the studied conditions or triggered

gelation by citrate ions.

In addition, the viscosity of concentrated BSA solutions obtained according to

magnetic nanoprobes ( DMS ) was estimated assuming the Stokes-Einstein relation for

the rotational diffusivity applies. The particle’s hydrodynamic diameter in BSA solutions

was assumed to be the same as those measured in water at different pH using DMS

measurements. The viscosity of the BSA solutions was calculated by fitting the Debye

model weighted by a lognormal size distribution to the out-of-phase DMS spectra. The

change in viscosity as a function of BSA concentration is seen in Figure 3-14 for each

buffer condition studied at pH = 2.7 and pH = 7.4. Like in bulk rheological

measurements, it was observed that the solution viscosity increases more rapidly at

lower pH values that at higher pH values. As we expected, Figures 3-13 and 3-14 show

a dependence between BSA concentration and increase in viscosity under the presence

of salt and buffers for both rheological measurements and DMS estimations. Overall, a

concentration increase is associated with a viscosity increase.

Page 93: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

93

Particularly, the results obtained via bulk rheology and DMS measurements

shown agreement in almost all the buffer conditions studied for BSA with the exception

for acidic conditions in citric acid + sodium citrate 0.1 M buffer without salt (A-B-NS

condition). Conversely, an order of magnitude difference was found between the

macromolecular viscosity and the viscosity experienced by the MNPs in their local

environment in this buffer condition as can be seen in Figure 3-15. This may be

explained by solid-like BSA gels formation. Hence, entangled protein filaments may be

restricting particle rotation, and eventually trapping the MNPs.

Furthermore, BSA conformation has been reported to transition from a closed so-

called “normal” (N) state to an open molecular structure (referred to as the “fast” state,

F) as pH decreases from neutral to acidic in the range of pH 7.4 to ~4.7-4 [174-177].

This transition is characterized by an opening of the molecular structure and

corresponding increase in protein dimensions, asymmetry, and increased exposure of

hydrophobic groups. As pH decreases from 4.7 to 2.7, BSA further unfolds to an

“expanded” (E) state, increasing in size and asymmetry [174, 175]. In addition, other

states have been observed, such as the “basic” (B) state seen at pH 8-10, which is

associated with structural fluctuations with a loosening of molecular structure and loss

of rigidity, and an “aged” (A) state at pH values greater than 10 [175, 176]. However,

this picture is not without controversy, as some studies claim that the most critical

change in volume and exposure of hydrophobic groups is during the N-F transition,

followed by a decrease in volume as pH is lowered to the E state [178].

The fact that at pH = 2.7 in citric acid + sodium citrate 0.1 M buffer without salts,

the solution viscosity becomes increasingly sensitive to increasing protein concentration

Page 94: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

94

in contrast with the other buffer conditions studied at acidic and basic scenarios could

be consistent with an expansion of BSA and/or with increased protein-protein repulsion

due to acquired charge. First, as the BSA molecule expands, a given mass of BSA

occupies a greater volume fraction in solution, thus resulting in greater suspension-

scale viscosity. Similarly, greater protein-protein repulsion would lead to a higher

effective volume fraction at a given mass concentration, again resulting in increased

viscosity. These two possible explanations for the pH dependence of the viscosity

observed using DMS measurements in Figure 3-15 are consistent with the above

discussion of changes in the charge and conformation of BSA by changing pH.

3.3.3 Rotational Diffusion and its Deviation from the Predictions Based on the Stokes-Einstein Relation

From the previous section, it was observed that viscosity values using DMS

evaluations under acidic conditions increased in comparison with those values at

neutral conditions. This is in part because of gelation of BSA due to protein

conformational changes as was discussed in previous sections, which were more

significant at buffered conditions with citrate ions. Figure 3-16 shows that the rotational

diffusion coefficient estimated using DMS measurements negatively deviate from the

Stokes-Einstein relation at acidic conditions. However, at neutral conditions (pH = 7.4)

the relation still applies because there is no conformational change in the BSA structure

(no BSA gel formation). The rotational diffusion predicted by the SE relation ( ,R SED ) was

estimated by using the bulk viscosity along with the lognormal size distribution at zero

protein concentration.

Our results agree with the observations made by Saraganpati et al. [161], who

concluded that protein-protein interactions increase in importance at high protein and

Page 95: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

95

salt concentrations. Hence, charge distribution and surface patchiness, relative

orientation, hydrophobic interactions, etc., potentially become relevant to describe the

viscosity of protein solutions.

According to reported DMS and DLS measurements, the particles do not

appreciably change in size or aggregate at the buffer conditions studied in the absence

of proteins except for pH = 2.7 buffered at salted conditions after 72 h. Thus, an

explanation for the tendencies observed in Figure 3-16, especially at acidic pH are not a

result of particle aggregation. In the case of particle aggregation this will result in

distortions in the out-of-phase susceptibility spectrum, which were not observed in all

the buffer condition evaluated in this study. Furthermore, because the rotational

diffusivity is inversely proportional to the hydrodynamic diameter, protein adsorption and

particle aggregation can count for the observed patterns in concentration and pH

dependence of the rotational diffusivity. However, since initial susceptibility values

remain mostly constant for all the buffered conditions studies increase in hydrodynamic

diameter distributions followed by particle aggregation and precipitation were discarded.

Therefore, a simpler justification could be that the observed changes in rotational

diffusivity are due to the dependence of the viscosity of the BSA solutions on BSA

concentration at pH value below and above the isoelectric point of BSA where BSA is

sensitive to conformational changes affecting the local environments of the MNPs.

3.4 Conclusions

The results reported herein demonstrate that under the buffer conditions studied,

the rotational diffusivity of magnetic nanoparticles in concentrated BSA protein solutions

follows the predictions of the Stokes-Einstein (SE) relation only at neutral buffer

conditions (pH = 7.4), which implies that the viscosity used is the concentration-

Page 96: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

96

dependent protein solution viscosity. However, when there is evidence of

conformational changes in BSA such as gelation patterns at acidic conditions, a

negative deviation from the rotational diffusion coefficient predicted by the SE relation

was observed. Our observations are important because they lend confidence to the use

of the Stokes-Einstein relation when predicting the dynamics of nanoparticles in protein

solutions at pH values above the isoelectric point of the protein itself mimicking

physiological conditions. These predictions serve to illustrate how magnetic nanoparticle

probes can be used to study particle mobility and colloidal stability in crowded protein

environments and to extract the mechanical properties of concentrated protein

solutions, which can be negligible via conventional bulk rheology measurements.

Because the reported method requires small volumes (< 200 µl) of sample, it is

potentially suitable for the study of solution containing difficult to obtain proteins (e.g.

monoclonal antibodies) and other biological fluids (e.g., mucus, synovial fluid, etc.).

In generalizing our observation of rotational diffusivity consistent with the values

obtained using the SE relation we point out that the particles used herein consisted of

cobalt ferrite clusters embedded in a covalently-bonded dense brush of poly (ethylene

glycol). We had previously reported how similar coated nanoparticles are colloidally

stable due to steric repulsion in a wide range of pH and in the presence of buffers and

proteins [167]. We therefore expect our observation of rotational diffusivity consistent

with the SE relation to be applicable to other similarly-stabilized inorganic particles

exhibiting Brownian relaxation and possessing dense polymer coatings, and which

prevent protein adsorption and/or nanoparticle aggregation in complex environments.

Page 97: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

97

Furthermore, the results reported in this chapter provide insight into the

concentration and pH dependence of concentrated protein solutions, particularly from

the perspective of a probe entity with the size in a nanoscale range. The observed

increase in sensitivity of solution viscosity with increasing protein concentration was

also consistent with other reports of expansion of BSA molecule with decreasing pH

below BSA isoelectric point.

Finally, it was observed that acid-cold gelation of BSA at acidic conditions was

triggered by the presence of sodium citrate ions, and the gelation pattern could be

reduced by increasing ionic strength in protein solutions using sodium chloride.

Page 98: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

98

Figure 3-1. Particle mobility and colloidal stability in concentrated protein solutions can be studied by means of monitoring magnetic particle rotation in an alternating magnetic field and calculating the rotational diffusivity of magnetic nanoparticles

Page 99: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

99

Figure 3-2. Thermogravimetric analysis of oleic acid coated cobalt ferrite nanoparticles and PEG-Silane coated MNPs (n = 3)

Page 100: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

100

Figure 3-3. Physical, magnetic, and colloidal characterization of the PEG-Silane coated cobalt ferrite nanoparticles. a) Representative TEM micrograph shows a physical diameter of ~14 nm (scale bar 100 nm). b) DMS spectra at pH = 7.4 in ultrapure deionized water showing characteristic Debye’s trend for thermally blocked MNPs. c) Size distributions for the inorganic core, obtained by TEM (blue line), and for the hydrodynamic diameter, determine from DLS (red line) and DMS measurements (black line)

Page 101: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

101

Figure 3-4. Stability of MNPs at pH values between 2.7 and 10 in water. a) Out-of-phase susceptibility data for PEG-cobalt ferrite MNPs. The peak position of the normalized out-of-phase susceptibility data does not vary significantly between changes in pH, indicating that the particle rotational dynamics is not modified by changes in pH. b) DLS measurements showing that particles remain polydisperse at different pH values, along with more common particle population remaining within the same hydrodynamic diameter

Page 102: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

102

Figure 3-5. Particle stability at the buffer conditions used to study PEG-cobalt ferrite MNPs mobility at higher BSA concentrations. a) Illustrative DMS spectra of MNPs at basic conditions, no buffered, and salted (137 mM NaCl) showing the characteristic Debye relaxation pattern. b) Illustrative DMS spectra of MNPs at basic conditions, buffered (PBS), and salted (137 mM NaCl) showing the characteristic Debye relaxation pattern. c) Out-of-phase susceptibility data for MNPs at the buffer conditions used to monitor MNPs mobility at high BSA concentrations, whereas the peak position of the normalized out-of-phase susceptibility data vary an order of magnitude between changes in pH in the presence of salts, indicating that the particle rotational dynamics is modified due to buffer and salt electrostatic interactions with protonated amine groups onto the surface of MNPs

Page 103: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

103

Figure 3-6. Equilibrium magnetization curves at 300 K for the PEG coated MNPs show a magnetization value about 100 A/m without mass normalization

Page 104: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

104

Figure 3-7. BSA protein solutions in acidic conditions at 4ºC. a) BSA at pH = 2.7, no buffered, no salted, BSA at pH = 2.7, no buffered, salted, BSA at pH = 2.7, buffered, no salted, and BSA at pH = 2.7, buffered, salted after continuous dialysis for 24 h. b)-e) BSA protein pH stability at different protein concentrations

Page 105: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

105

Figure 3-8. BSA solutions showing aggregates/cluster and gel formation. a)-d) BSA in acidic conditions before (left) and after (righ) filtration using a 0.22 um syringe filter. e) BSA in citric acid + sodium citrate 0.1M salted 137 mM NaCl supernatant after protein concentration and posterior centrifugation to remove protein gel structures. f) BSA gel bottom after protein concentration and posterior centrifugation. g) BSA in basic conditions buffered in PBS (pH = 7.4) before (left) and after (right) filtration using a 0.22 um syringe filter. No aggregation was observed before and after dialysis, and gelation was observed after protein concentration

Page 106: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

106

Figure 3-9. Decreasing concentration of BSA at pH = 2.7 with an ionic strength 137 mM NaCl (sodium citrate + citric acid 0.1 M buffer) in the AC-susceptometer vials (200 µL). The top image corresponds to protein suspension without MNPs, and bottom image corresponds to protein suspension with PEG coated MNPs after 48 h of exposure

Page 107: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

107

Figure 3-10. Normalized out-of-phase dynamic magnetic susceptibility for nanoparticles in BSA at pH = 2.7 buffered using citric acid + sodium citrate 0.1 M. 137 mM NaCl

Page 108: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

108

Figure 3-11. Rotational diffusivity of PEG coated MNPs in acidic and basic conditions with and without the presence of salts. D0 is the rotational diffusivity of PEG coated MNPs in buffer without added proteins

Page 109: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

109

Figure 3-12. Bulk rheological measurements at pH = 2.7 buffered and no salted (A-B-NS conditions)

Page 110: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

110

Figure 3-13. Viscosity measurements of BSA in buffered and salted conditions at pH = 2.7 and pH = 7.4 obtained via bulk rheological measurements

Page 111: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

111

Figure 3-14. Viscosity measurements of BSA in buffered and salted conditions at pH = 2.7 and pH = 7.4 obtained via DMS measurements

Page 112: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

112

Figure 3-15. Comparison between viscosity values of BSA protein solutions in buffered and salted conditions at pH = 2.7 and pH = 7.4 obtained via macromolecular rheology and DMS measurements

Page 113: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

113

Figure 3-16. Deviation of the Stokes-Einstein relation in BSA solutions below and above the BSA isoelectric point

Page 114: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

114

CHAPTER 4 DETERMINING THE MOBILITY OF POLYMER COATED NANOPARTICLES IN

BLOOD AND TUMOR TISSUE USING MAGNETIC RELAXATION MEASUREMENTS

4.1 Introduction

Depending on the inorganic metal core (e.g. iron oxide or cobalt ferrite), magnetic

nanoparticles may possess either a freely rotating magnetic dipole (Néel relaxation

mechanism) or a dipole fixed along a crystal direction (Brownian relaxation mechanism).

These relaxation mechanisms represent the theoretical foundation to understand how

the response of magnetic nanoparticles subjected to alternating magnetic fields can

provide useful information in terms of nanoparticle-protein interactions that lead to

adsorption of proteins or aggregation of particles in complex biological environments. In

this work, we tested a method to determine hydrodynamic diameter distributions of

magnetic nanoparticle suspensions in blood and tumor tissue by using the Debye model

applied to dynamic magnetic susceptibility measurements to assess nanoparticle-

protein interactions and colloidal stability prior in vivo clearance experiments. These

colloidal stability measurements were incorporated as a complementary technique to

predict clearance of magnetic nanoparticles after intravenous injection in rodent models

and provide a systematic route to design long term circulation magnetic nanoparticles

for biomedical applications.

4.2 Materials and Methods

DMEM-high glucose power media (D5648, Sigma-Aldrich), fetal bovine serum

(F6178, Sigma-Aldrich), L-asparagine (A4159, Sigma-Aldrich), 10% L-Glutamine

(25030081, Thermo Fisher), penicillin-streptomycin (15140122, Thermo Fisher), 175

cm2 culture flasks (431466, CorningTM), Lithium-Heparin tubes (365965, BD).

Page 115: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

115

Iron (III) chloride hexahydrate (FeCl3·6H2O, 97%, ACS reagent), cobalt (II)

chloride hexahydrate (CoCl2·6H2O, 98%), oleic acid, mono-methoxy PEG MW 2000,

sodium hydroxide (NaOH, ≥98%) pellets, nitric acid (HNO3, 70%), iron (III) nitrate

nonahydrate (≥98%), dimethyl sulfoxide (DMSO, ≥98%), N-hydroxysuccinimide (NHS),

polyethyleimine, branched average Mw ~25,000 by LS and CM-Dextran sodium salt

Ave. MW 10,000-20,000 were purchased from Sigma Aldrich. Tetrametylammonium

hydroxide ((CH3)4NOH, 10%), 1-Ethyl-3-(3-dimethylaminopropyl) carbodiimide

hydrochloride (EDC), sulfo-N-hydroxysuccinimide (Sulfo-NHS), sulfo-N-

hydroxysuccinimide-acetate (sulfo-NHS-acetate), and CBQCA protein quantitation kit

were purchased from Thermo Fisher Scientific. 3-aminopropyltriethoxysilane

(NH2(CH2)Si(OC2H5)3) (APS) was purchased from TCI America. Bionized NanoFerrite

particles with nominal sizes of 80nm and 100nm (BNF-Starch MNPs) were purchased

from Micromod Partikeltechnologie GmbH and used as received. 2µm nylon syringe

filters and 100 kDa centrifugal filters (Amicon) were purchased from Sigma-Aldrich.

4.2.1 Animals and Tumor Model

A subline of MDA-MB-231 cells [179] (a gift by Dr. Dietmar W. Siemann,

Department of Radiation Oncology, University of Florida, College of Medicine) were

cultured in vitro in DMEM-high glucose power media supplemented with 10% fetal

bovine serum, 2.5 wt. % L-asparagine, 10% L-Glutamine and 10% penicillin-

streptomycin. Cells were cultured and incubated at 37°C in a 5% CO2 atmosphere. A

cell suspension of ~6 x 106 MDA-MB-231 cells in 200 µL phosphate buffered saline was

injected subcutaneously into the abdominal fat pad of 6-week old female NOD SCID

mice (Envigo). After six weeks post cell inoculation, a tumor ~ 1000 mm3 in volume was

formed. Tumor growth was monitored every three days using caliper measurements.

Page 116: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

116

Animal handling and procedures were approved and performed according to the

guidelines of the Institutional Animal Care and Use Committee of The University of

Florida.

4.2.2 Synthesis of Cobalt Ferrite Magnetic Nanoparticles

Thermally blocked cobalt ferrite magnetic nanoparticles were synthetized by

thermal decomposition of an organometallic precursor as described by Herrera et.

al.[180] and by the co-precipitation method by controlling the flow rate addition of metal

ions in sodium hydroxide under boiling conditions,[148] respectively.

4.2.2.1 PEI coated cobalt ferrite MNPs

Oleic acid coated cobalt ferrite MNPs obtained by thermal decomposition were

used for subsequent aqueous phase transfer mediated by the cleavage of the double

bond in the oleic acid to form a carboxylic terminal [181]. Then surface modification of

carboxylated MNPs was carried out as described by Cruz-Acuña et. al.[182] with some

modifications. In acidic conditions, coupling with branched 25 kDa PEI via EDC/ sulfo-

NHS chemistry was performed to obtain PEI 25 kDa coated cobalt ferrite (tICO-PEI-

25kDa) MNPs. Briefly, 4 mg EDC and 0.5 mg sulfo-NHS were dissolved with 5 mL of

carboxylated MNPs at pH 5 (10 wt. % in deionized water) and left to react 10 min for

carboxylic activation. Then, a solution of branched PEI in deionized water (20 wt. % 25

kDa MW, pH 4.5-5) was mixed with the activated carboxylic MNPs and left to react

under a high-intensity ultrasonic processor (XL2020, Misonix Inc.) to break aggregates

(1 h in an ice bath). After reaction, MNPs were dialyzed into ultrapure deionized water

using 100 kDa centrifugal filters (Millipore) centrifuged at 5000 rcf for 10 min at 4°C

three times.

Page 117: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

117

4.2.2.2 CMDx coated cobalt ferrite MNPs and PEG coated cobalt ferrite MNPs

The cobalt ferrite particles synthetized via co-precipitation[148] were surface

modified with HNO3 2 M and hydrothermally treated with ferric nitrate according to

Gomes et al.[149]. Then MNPs were peptized using tetramethylammonium hydroxide

(TMAO) for further surface modification using either condensation of 3-aminopropyl

triethoxysilane (APS) molecules or oleic acid chemisorption i) to covalently attach

CMDx or ii) to perform a ligand exchange with PEG-Silane.

Bohorquez et. al. [78] described the procedure to coat cobalt ferrite MNPs

synthetized by the co-precipitation method with CMDx after APS condensation. To coat

cobalt ferrite nanoparticles with oleic acid, an adsorption reaction was carried out as

described by Hubbard et. al.[160] Thus, ligand exchange with PEG-Silane MW 2 kDa

was performed as described by Barrera et. al.[70] using as-synthetized oleic acid

coated cobalt ferrite nanoparticles suspended in toluene. To remove free polymer

excess, polymer coated MNPs were dialyzed against deionized water using 100 KDa

centrifugal filters (Millipore) centrifuged at 5000 rcf for 10 min at 4°C three times.

4.2.3 Iron Oxide Magnetic Nanoparticles

Iron oxide MNPs exhibiting a combination of Brownian and Néel relaxation

mechanisms were synthetized by chemical co-precipitation of ferric and ferrous salts in

alkaline medium and using TMAO as a peptizing agent.[183] CMDx and PEG coatings

were grafted onto the surface of the iron oxide nanoparticles with slightly modifications

to the described procedures developed by Herrera et. al[150] and Barrera et. al[70] for

iron oxide MNPs. A ligand exchange with PEG-Silane MW 2kDa was performed with

iron oxide cores modified with oleic acid via adsorption at 80°C as was described for

cobalt ferrite MNPs synthetized by using the co-precipitation method.

Page 118: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

118

4.2.4 Surface Modification, Amine Functionalities and Amine Quantification

4.2.4.1 PEG coated MNPs – amine functionalities

PEG-Silane polymers synthetized to obtain colloidally stable PEG coated MNPs,

contain some traces of APS molecules with free amine groups. Then, after ligand

exchange PEG particles possess exposed amine groups that can vary from batch to

batch for a fixed PEG molecular weight. Previously, it has been demonstrated that for

low molecular weight PEG (0.75 kDa - 1 kDa), amine groups are exposed contributing

to an overall positive surface charge whereas particle coated with higher molecular

weight PEG were almost neutrally charged (up to 5 kDa).[167] Therefore, a molecular

weight of 2 kDa was selected for particle mobility studies in whole blood and tumor

tissue using DMS.

4.2.4.2 CMDx coated MNPs and capped CMDx coated MNPs – amine functionalities

EDC-mediated coupling reaction between amine groups in APS coated MNPs

and CMDx polymer chains proceeds with low conjugation efficiency, then unreacted

amine groups can remain covered by the CMDx polymer shell after reaction.

Domenech et al.[184] were able to conjugate the dye BODIPY-FL to the free amine

groups in cIO-CMDx via EDC/NHS reaction, suggesting the presence of unreacted

amine groups. For DMS measurements, free amine groups in cIO-CMDx MNPs were

acetylated with sulfosuccinimidyl acetate (sulfo-NHS-acetate) to irreversible cover

primary amine groups and avoid undesirable particle aggregation due to electro-viscous

phenomena mediated between a polyelectrolyte particles (particle with both amino and

carboxylic functionalities) in PBS, blood, and tumor tissue. To remove the free polymer

excess and unreacted EDC/NHS, CMDx coated MNPs were dialyzed against deionized

Page 119: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

119

water using 100 KDa centrifugal filters (Millipore) centrifuged at 5000 rcf for 10 min at

4°C three times. For Capped CMDx coated MNPs, MNPs were loaded into MWCO 10

kDa dialysis cassettes (Slide-A-LyzerTM, Thermo Fisher) after acetylation reaction and

dialyzed against PBS under mild agitation at 100 rpm for 24h.

4.2.4.3 Quantification of amine groups in the surface of the magnetic nanoparticles

The number of amine groups per particle was estimated using 3-(4-

carboxybenzoyl) quinolone-2-carboxaldelhyde (CBQCA) which reacts to primary amines

in the presence of thiols or cyanide at basic conditions. The fluorescent product of

CBQCA has an absorption peak excitable at 430-490 nm with maximum emission at

around 560 nm. NTA measurements were used to determine particle concentration per

each particle stock solution. First, the linear dynamic range of CBQCA response was

studied under three fixed peptized iron oxide particle suspensions (1.69 mgFe/mL, 1

mgFe/mL and 0.25 mgFe/mL) using ethanolamine as calibration curve standard. Then,

a post validation of particle interference under a fixed ethanolamine concentration in

CBQCA assay was performed in a particle concentration range between 0-2 mgFe/mL.

From these conditions the optimal fluorescence signal vs particle concentration was

0.25 mgFe/mL. This is in part because the dynamic range was the highest between 0-

1.5 µg ethanolamine increasing assay sensitivity. Afterwards, CBQCA unknown

samples (500 µL) were prepared from particles stocks in water diluted up to 0.25 mg

Fe/ml with 0.1 M sodium borate at pH = 9.3. Then, the derivatization reaction was

carried out by mixing 130 µL of CBQCA unknown sample with 5 µL of 20 mM KCN and

10 µL of a 4 mM CBQCA working solution for 2 hours in a 96-well plate microplate for

analysis in a fluorescence microplate reader. Amine groups were quantified using an

Page 120: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

120

ethanolamine calibration curve prepared under the same previous reaction conditions

as follow, 130 µL of peptized iron oxide MNPs in 0.1 M sodium borate at pH = 9.3 (0.25

mgFe/mL) were mixed with 1 µL ethanolamine standard solutions, 5 µL of 20 mM KCN

and 10 µL of a 4 mM CBQCA working solution. Amine quantification was performed by

triplicates per particle used in this study.

4.2.5 Characterization of Magnetic Nanoparticles

Infrared spectroscopy in a Varian 800 Transform Infrared (FTIR) equipment was

used to study changes in functional groups obtained in the synthesized coated MNPs.

Powder samples were placed on a Pike attenuated total reflectance (ATR) stage with

ZnSe window. The results are presented in a plot of transmittance (dimensionless units)

vs. wavenumber (cm-1).

A Hitachi 7600 transmission electron microscope (Hitachi High Technologies

America, Pleasanton, CA) equipped with a MacroFire slow-scan CCD camera

(Optronics, Goleta, CA) and AMT Image capture software (Advanced Microscopy

Techniques, Danvers, MA) operating at 120 kV acceleration voltage was used to image

the nanoparticles. Stock samples at 0.5 mgFe/mL (5 µL) were applied onto a ultrathin

Formvar-carbon-coated nickel grid and air dried. The physical diameter of the

nanoparticles has been determined from the TEM pictures using ImageJ and counting

at least 200 particles. The histogram of particle size distribution was fitted to a

lognormal size distribution, to estimate the number-weighted mean diameter and the

geometric deviation of the MNPs.

The hydrodynamic size distributions of the synthetized MNPs were determined

by dynamic light scattering (DLS) and dynamic magnetic susceptibility (DMS)

measurements in water with adjusted ionic strength using 1 mM KNO3 at pH 7.4 at

Page 121: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

121

25°C. DLS and zeta potential measurements were performed using a particle analyzer

(Zeta PALS, Brookhaven Instruments). The nanoparticles stock solutions were prepared

in water and then filtered through 0.2 µm nylon syringe filter twice prior to

measurements and then diluted to ~0.1 mgFe/mL. Volume-weighted size distributions

were fitted to a lognormal distribution and the arithmetic mean diameter and standard

deviation were calculated along with zeta potential measurements (10 runs per sample).

DMS measurements using a commercial AC susceptometer (Dynomag system,

Acreo Swedish ICT) were performed at room temperature. During measurements, the

amplitude of the excitation field varied between 0-0.5 mT, and the frequency interval

was evaluated in a range of 10 Hz to 160 kHz. All DMS measurements were fitted to the

generalized Debye’s model assuming that the particles follow a lognormal size

distribution and the viscosity of the fluid carrier assumed as water (0.001 Pa.s).

Nanoparticle tracking analysis (NTA) was performed using a Nanosight LM-10

system (Malvern Intruments) with 405 nm laser excitation. Particle stock solutions in

water were diluted in KCl 0.1 M by a 106 dilution factor and filtered through 0.2 µm nylon

syringe filter. Samples for analysis were injected into the sample chamber using a

sterile syringe, and video capture was initiated immediately. All measurements were

collected at room temperature with the camera level to the maximum value. For each

measurement, three videos were taken with a video acquisition time auto defined and

analyzed (50-100 particles per frame, 108-109 particles per mL) using NTA 3.1 software.

Number-weighted size distributions were fitted to a lognormal distribution and converted

to volume-weighted size.

Page 122: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

122

For iron oxide particle stocks in water, the iron content was determined using a

colorimetric assay based on the complexation of Fe+2 with 1,10 phenanthroline. Ten

microliters of particle stocks in water were digested with 1 mL of 70% nitric acid (Optima

Grade, Fisher Scientific) in a dry bath at 101°C overnight. Then, 10 µL of digested

sample were evaporated at 115°C for 1 hour, followed by addition of 46 µL of deionized

water and 30 µL of hydroxylamine hydrochloride (8 M). After 1 hour of reaction,

reduction from Fe+3 to Fe+2 was completed and 49 µL of sodium acetate (1.2 M) were

added, followed by 75 µL of 1,10-phenanthroline monohydrate (13 mM). Afterwards,

100 µL of each sample were transferred into a 96-well quartz microplate for absorbance

analysis in a microplate reader at 508 nm. Absorbance data was used to determine the

iron concentration against an iron standard calibration curve. Samples and calibration

standards were run in triplicate.

The magnetic core size and saturation magnetization of MNPs stocks in water

(~0.5 mgFe/mL) were determined by measuring the response of the equilibrium

magnetization under the application of a DC field at 300 K in a magnetic range from 6 to

-6T using a Quantum Design MPMS-3 Superconducting Quantum Interference Device

(SQUID) magnetometer. The Langevin equation weighted using a lognormal size

distribution was used to calculate the magnetic diameter of the nanoparticles and the

saturation magnetization was obtained by averaging 10 points in the maximum

saturation region and normalized by iron quantification.

4.2.6 Blood Collection and Tumor Harvesting

Mice were randomly selected for particle mobility assessments when tumors

reached a size between 800-1000 mm3 in volume. Blood was collected by cardiac

puncture using sodium citrate-wetted syringes (0.1 v/v %) and 25G needles and

Page 123: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

123

transferred into capillary blood collection tubes. Then blood collection tubes were placed

in a tube rotator with slow rolling wrist motion to avoid clot formation and hemolysis.

Whole blood was used immediately to perform DMS measurements. Tumors were

harvested under sterile conditions, kept on ice and used the same collection day.

Processing consisted of transferring the tumor into a Petri dish containing PBS and

cutting fragments using a disposable biopsy punch (10-12 fragments of 6 mm per

tumor) to: i) assess colloidal stability in tumor tissue using DMS measurements, ii)

perform Prussian blue staining, and iii) nanoparticle-tumor tissue visualization using

transmission electron microscopy (TEM).

4.2.7 Assessment of Particle Mobility Using Dynamic Magnetic Susceptibility Measurements

Colloidal stability as a function of particle mobility was assessed using a

commercial AC susceptometer (Dynomag system, Acreo Swedish ICT) at room

temperature. The amplitude of the excitation field varied between 0-0.5 mT, and the

frequency interval was evaluated in a range of 10 Hz to 160 kHz. A representation

scheme for the sample preparation during particle mobility assessments is shown in

Figure 4-1. Blood and harvested tumor tissue hydrated with PBS were used

immediately. Briefly, 10 µl of polymer coated MNPs (~39 µgFe) suspended in PBS were

mixed gentle with 90 µL of whole blood and immediately tested in the AC

susceptometer (t = 0 h). For tumor tissue sample evaluation, a tumor fragment was

placed in a polycarbonate capsule size 4 and a hole was opened in the closed capsule

using a 26G needle. Then 10 µL of polymer coated MNPs (~39 µgFe) suspended in

PBS were loaded in a 25 µL Hamilton syringe coupled with a Hamilton needle 26G

0.75” 12° bevel and slowly injected in the tumor fragment placed inside the

Page 124: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

124

polycarbonate capsule. After infusion needle was left inside the tissue for 5 min post-

injection to optimize nanoparticle distribution and then removed carefully. The tissue

fragment was immediately tested in the AC susceptometer (t = 0 h). Next, whole blood

and tumor tissue samples were measured at t = 24 h as a second evaluation time point.

One mouse was used per replicate. Three replicate were evaluated per condition.

Dynamic magnetic susceptibility data was fitted to the generalized Debye’s

model assuming that the particles follow a lognormal size distribution and the viscosity

of water (0.001 Pa.s), obtaining as the fit parameters the lognormal diameter distribution

and geometric deviation which represents a measure of the particle mobility in PBS,

mice blood and tumor tissue. Along with the obtained diameter distributions, initial

susceptibility values of MNPs in mice blood and tumor-tissue were normalized with

initial susceptibility values in PBS at t = 0 h, and the arithmetic mean diameter and

standard deviation were estimated to evaluate particle polydispersity in biological

millennium as a function of time.

4.2.8 Tissue Preparation for Electron Microscopy

Tissue processing and embedding into Epon resin for examination was

performed as follows: harvested tumor tissue fragments injected with MNPs (~1 mm3)

were fixed in a mixture of 4% formaldehyde – 2% glutaraldehyde in PBS overnight

(primary fixation). Afterwards, samples were washed with PBS in order to remove

fixation cocktail and a post fixation treatment with 1% osmium tetraoxide/ 2-

mercaptoethanol (2-ME) in 0.1M Na cacodylate was performed. Samples were washed,

dehydrated through a graded series of ethanol, and embedded with Quetol 651 resin

following procedures that have been previously published [185]. Ultrathin sections of

samples (70 nm) containing well-preserved ducts were mounted on uncoated copper

Page 125: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

125

grids. Sections with and without uranyl acetate (alcoholic 8%)/ Venable’s lead citrate

stains were examined using a Hitachi 7600 transmission electron microscope (Hitachi

High Technologies America, Pleasanton, CA) equipped with a MacroFire slow-scan

CCD camera (Optronics, Goleta, CA) and AMT Image capture software (Advanced

Microscopy Techniques, Danvers, MA) operating at 120 kV acceleration voltage.

4.3 Results and discussion

The particle coatings studied herein were selected due to their relevance in

several biomedical applications. First, a non-viral gene transfection mediator, branched

polyethyleneimine (b-PEI), was studied for being a polycation with a large buffering

capacity at physiological conditions [186]. PEI coated MNPs have been used in

magnetofection to improve DNA transfection efficiency [187] and for targeted gene

delivery [188, 189]. The second selected coating was carboxymethyl dextran (CMDx).

CMDx coated MNPs have been used extensively for magnetic cell labeling [190],

magnetic resonance imaging[191] and hyperthermia studies for cancer treatment [16,

184, 192-194]. They are well known to have non-specific interactions in biological

systems due to overall negative surface charge, phagocytic cell uptake, and

biodegradability. As third coating alternative, we selected hydroxyethyl starch (BNF-

Starch). This is a non-ionic starch derivative biodegradable stabilizer in BNF-Starch

commercial MNPs used mostly for magnetic fluid hyperthermia applications [195].

Finally, as a fourth coating we chose polyethylene glycol (PEG), which provides immune

evasion and opsonization resistance [139] for magnetic nanoparticles in biological

environments, a suitable option for targeted therapeutics. Other surface coatings or

modifications included as reference were amino-propyl-triethoxysilane (APS)

condensation and sulfosuccinimidyl acetate (sulfo-NHS-acetate) induced acetylation in

Page 126: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

126

CMDx coated MNPs. APS was studied due to its highly positive surface charge, and

resistance to enzymatic attack in lysosomes used in earlier magnetic fluid hyperthermia

studies [196], and significance as an intermediate step in preparation of covalently

bound CMDx coated MNPs [73]. To determine the impact of controlled surface

chemistry after CMDx immobilization onto APS coated MNPs, unreactive primary

amines were acetylated (capped CMDx coated MNPs) and particle mobility tested.

Acetylation of primary amine groups performed with sulfosuccinimidyl acetate forms an

amide bond under basic conditions (R-NH3+ groups are converted to R-NH-CO-CH3). It

has been shown that partial acetylation of lysine residues reduce intraprotein

crosslinking/aggregation [197] and we tested this approach to cover unreactive primary

amine groups in cIO-CMDx MNPs that could interfere with particle stability in biological

systems.

The FTIR spectra of five representative coated MNPs used here are presented in

Figure 4-2. For the PEG coated nanoparticles, the Si-O-Fe and Fe-OH peaks were

detected between 800-900 cm−1. The peak at 1095 cm−1 is characteristic of –COC

vibrations from the ether group in PEGSilane. Also, a strong band at 2780 cm−1 was

attributed to C-H stretching vibrations. A band appeared at 1635 cm−1, characteristic of

the –C(=O)–NH vibration in PEG-Silane [198, 199].

Silanized MNPs formed small particle aggregates as a result of co-precipitation

synthesis and APS condensation which lead to the formation of nanoclusters embedded

in siloxane shell (Figure 4-3). Table 4-1 shows a summary of the physicochemical

characterization of the MNPs in water at pH ~7.4 containing 1 mM potassium nitrate

(KNO3) or 1 mM potassium chloride (KCl) to control ionic strength. PEI coated MNPs

Page 127: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

127

and APS coated MNPs showed the expected positive surface through zeta potential

measurements. Differences in hydrodynamic diameter distributions for both particles are

due to the employed synthesis method and surface modification procedure. The cationic

PEI coated MNPs made by the thermal decomposition method were stabilized via

electrostatic repulsion and steric hindrance, showing a narrow size distribution in

comparison with APS grafted MNPs.

Table 4-1. Physicochemical characterization of coated magnetic nanoparticles used in this study.

Sample Inorganic core

Coating ɀ- potential, [mV]

DH (DLS),

[nm]

DH (NTA),

[nm] DH (DMS),

[nm]

tICO-PEI Cobalt ferrite

b-PEI 25kDa 23 ± 1 37 ± 16 64 ± 22 32 ± 9

cICO-CMDx Cobalt ferrite

CMDx 10 kDa

-30 ± 2 72 ± 24 81 ± 31 79 ± 23

cICO-PEG Cobalt ferrite

PEG-Silane 2 kDa

0 ± 1 52 ± 20 171 ± 81 54 ± 16

BNF-Starch 80 nm

Iron oxide

Hydroxyethyl starch

-4 ± 1 103 ± 24 - 102 ± 23

BNF-Starch 100 nm

Iron oxide

Hydroxyethyl starch

-6 ± 1 121 ± 38 - 129 ± 34

cIO-APS Iron oxide

APS 34 ± 2 69 ± 28 139 ± 63 89 ± 52

cIO-CMDx Iron oxide

CMDx 10 kDa

- 38 ± 1 73 ± 22 64 ± 22 68 ± 39

Capped cIO-CMDx

Iron oxide

CMDx 10 kDa

- 16 ± 1 70 ± 26 64 ± 21 64 ± 32

cIO-PEG Iron oxide

PEG-Silane 2 kDa

4 ± 0 59 ± 31 59 ± 19 49 ± 16

All CMDx coated MNPs presented a negative surface charge. However, capped

cIO-CMDx particles showed a reduction in the absolute zeta potential value (from - 38 ±

1 mV to - 16 ± 1), which confirms that acetylation modified the ɀ-potential in the

interfacial double layer. Due to amide bonds formation after acetylation reaction, amide

Page 128: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

128

groups should not contribute with a positive surface charge at physiological conditions,

thus no signal by using the CQBCA assay. This change in ɀ-potential could may be

attributed to electrostatic screening effect after dialysis against phosphate buffered

saline (PBS) for capped cIO-CMDx MNPs. Increase in salt content leads to lower zeta

potential values due to the increased impact of electrostatic shielding. Differences in

hydrodynamic size distributions were not noticed between cIO-CMDx and capped cIO-

CMDx MNPs based on optical and magnetic measurements.

The ɀ-potential measurements for BNF-Starch MNPs remained almost neutral at

the studied pH and ionic strength conditions. Size distribution estimations via DLS and

DMS measurements were in good agreement. PEG coated MNPs showed a minimal ɀ-

potential value which confirms that steric stabilization is predominant in this colloidal

system.

Nanoparticle tracking analysis (NTA) measurements showed a slight increase in

hydrodynamic size distributions in comparison with DLS and DMS analysis for most of

the particles used in this study with the exception of positively charged MNPs. This is in

part because, NTA can measure single particles and absolute particle concentration

thus it is more prone to detect a wider range of particle sizes per sample. For cationic

particles, the deviation was more dramatic on converting number-weighted distributions

to volume-weighted distributions. DMS measurements confirmed that particles present

relax predominantly by the Brownian relaxation mechanism exhibiting a well-defined

susceptibility spectrum that follows the characteristic Debye’s model.

To confirm the presence of free amine groups, shielded by the longer polymer

chains of CMDx and PEG, but which may still cause nanoparticle-blood or nanoparticle-

Page 129: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

129

tissue interactions, a chromatographic derivatization reagent was used to quantify

accessible primary amines in solution. Figure 4-4 shows a comparison between a) zeta

potential measurements and b) amine quantification results using 3-(4-carboxybenzoyl)

quinolone-2-carboxaldehyse (CBQCA) for the particles used in this study. Number of

particles per particle stock solution were estimated using NTA measurements.

Amine quantification was carried out under basic conditions (pH ~9.3). The

reaction occurs in presence of potassium cyanide (KCN), starting with a nucleophilic

substitution between the aldehyde terminal in CBQCA and the cyanide ion ( ͞ :C≡N).

Then primary amino groups react with this intermediate product to form a highly

fluorescent isoindole derivative [200, 201]. Important features of CBQCA as a labeling

agent are that i) it does not fluoresce unless it reacts with primary amino groups (excess

of CBQCA does not need to be removed from sample) and ii) the stability of the

derivatization reaction [202].

Branched PEI has primary, secondary and tertiary amino groups in a ratio of

1:2:1, and up to 25% of these amino groups are protonated under physiological

conditions[203]. However, the primary amine groups in b-PEI were covalently coupled

to the carboxylic groups onto MNPs using EDC and sulfo-NHS. Then, primary amine

groups did not contribute for the overall ɀ-potential value obtained, and they were

unable to be detected via CBQCA as shown Figure 4-4a and 4-4b.

For cICO-CMDx MNPs and cICO-PEG MNPs we calculated 49 ± 2 amine

groups/MNP and 296 ± 67 amine groups/MNP, respectively. No amine groups were

expected in BNF-Starch MNPs, as was confirmed via CBQCA detection. The highest

value obtained was 458 ± 83 amine groups/MNP for cIO-APS MNPs as expected, then

Page 130: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

130

after CMDx coupling the number of free amine groups was reduced to 167 ± 5 amine

groups/ MNP which signifies that 64% of the free amine groups in the APS coated

MNPs were utilized at the reaction conditions employed for CMDx immobilization. In

order to control the remaining free amine groups, the acetylation reaction via sulfo-NHS-

acetate modification for iron oxide MNPs showed a capped efficiency of 100% with no

amine signal contribution. Finally, for cIO-PEG MNPs it was calculated 166 ± 17 amine

groups/MNP.

Equilibrium magnetization measurements at 300 K in aqueous solutions

demonstrated that all MNPs exhibited superparamagnetism. As a representative

equilibrium magnetization curve, Figure 4-5 shows the magnetic response of PEI coated

MNPs. By fitting the data to the Langevin function using the procedure established by

Chantrell [169], the magnetic diameter was 11 ± 4, while the volume fraction ( ) of the

dilute stock solution was 0.03%.

4.3.1 Particle Mobility in PBS, Whole Blood and Tumor Tissue Using Magnetic Relaxation Measurements

MNPs in PBS, blood and tumor tissue were subjected to DMS measurements to

track particle mobility. The following graphs show the particle mobility patterns obtained

for i) polymer coated cobalt ferrite MNPs (Figures 4-6 and 4-7), ii) thermally blocked

polymer coated iron oxide MNPs (Figures 4-8 and 4-9), and iii) polymer coated iron

oxide nanoparticles with mixed relaxation (Figure 4-10 and 4-11). MNPs predominantly

followed the behavior of the Debye model. However, iron oxide nanoparticles as-

synthetized by co-precipitation method shows a mixture of magnetic relaxation

components [157]. DMS spectra evaluating changes in peak frequency from the out-of-

phase susceptibility have provided sufficient information to understand what was

Page 131: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

131

happening in the local environment of the MNPs upon contact with biological milieu for

24 hours.

In Figure 4-6a we observed that upon contact with tumor tissue PEI MNPs

deviate from the Debye model and the Brownian peak almost disappeared may be due

to possible particle aggregation and/or posterior particle precipitation. In blood some

particle mobility still remains but it is limited in comparison with particle mobility in PBS.

Hydrodynamic diameter distributions obtained via DMS susceptibility measurements

revealed that particle polydispersity increased in tumor tissue environments (Figure 4-

6b). We can attribute this to electrostatic interactions of particle with blood proteins

playing a significant role in particle aggregation. Relative retardation of particle mobility

due to electrostatic interactions has been observed with cationic particles suspended in

hyaluronic acid solutions and bovine vitreous humor ex-vivo using single particle

tracking (SPT) microscopy analysis [204]. Positively charged particles have shown a

heterogeneous diffusion pattern in vitreous humor, where a small fraction remains

mobile and the majority of particles immobilized were attached to collagen fibrils in the

vitreous humor, which may be a close representation of what is happening in our tumor

tissue experiments.

Figure 4-7 shows that PEI coated particles dramatically lose their magnetic

response after 24h of incubation in mouse blood and tumor-tissue (normalized initial

susceptibility values decreased to 0.2). Electrostatic interactions seem to determine

particle aggregation in this system and we cannot involve other interaction component

such as hydrophobicity with the information collected for this experimental design. This

Page 132: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

132

could happen due to negative charge contribution of proteins in blood and several

fibrous connective tissues in tumor stroma interacting with positively charged MNPs.

In contrast, CMDx and PEG coated nanoparticle experienced mobility up to 24h

of evaluation in PBS and blood because of no reduction in the normalized initial

susceptibility values, which mean that particles remained stable in size responding by

physical rotation to an applied external magnetic field. In contrast, in tumor-tissue

explants differences were observed in particle mobility between CMDx and PEG

coatings. Cobalt ferrite coated CMDx MNPs slightly aggregate when interacting with

tumor-tissue fragments (t = 0h), and after 24h the out-of-phase Brownian peak shifts to

a lower frequency range as a result of an increase of particle aggregation and/or particle

immobilization upon particle contact with tumor tissue explants.

BNF-Starch 80nm MNPs remain stable in blood and tumor tissue in comparison

with BNF-Starch 100 nm MNPs (Figure 4-8). Particle mobility restriction close to 90%

was observed for BNF-Starch 100nm MNPs in tumor tissue after 24h (Figure 4-9). BNF-

Starch 80nm and CMDx MNPs showed partial mobility in blood and tumor tissue after

24h. However, BNF-Starch 100nm MNPs have shown lesser magnetic response after

24h according to the dynamic susceptibility spectra evaluation. An explanation of this

could be particle diffusion inside the tumor matrix. Since BNF-Starch 100 nm particles

are bigger particles and less colloidally stable than CMDx and CMDx-Blocked MNPs,

they diffuse slowly until coating desorbs or clusters get internalized by cells still alive in

the time window of this study. CMDx-blocked nanoparticles even though are smaller

and diffuse faster, their specific surface modification allows them to not attach in the cell

matrix, helping them diffuse freely. If there was the case of MNPs internalization in cell

Page 133: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

133

compartments during the time window studied, in less than 24h particles could be

already discarded to the extracellular space maintaining a hydrodynamic size less than

100 nm, which can be related to observe a well-defined Brownian component or a

constant initial susceptibility value measured using the AC susceptometry.

Iron oxide CMDx coated nanoparticles due to free amine groups became highly

polydisperse after CMDx conjugation (Figure 4-10). However, this can be tuned by

blocking amine groups using sulfo-NHS acetate, an inert molecule. This degree of

polydispersity is detectable by optical measurements such as DLS and NTA but

becomes more relevant using dynamic magnetic susceptibility measurements in order

to study colloidal stability in blood and tumor tissue due to the complexity of these

biological environments. These findings demonstrate that surface modification needs to

be monitored and improved specifically in targeting strategies and sensing applications

to avoid confusing the electrostatic interactions with particles in bio-environments

instead of targeting-ligand interactions.

Since this set of iron oxide nanoparticles was synthesized by using the co-

precipitation method, their magnetic relaxation mechanism is a combination between

Brownian and Néel relaxation mechanisms, hence we cannot rely on the analysis of the

normalized initial susceptibility values with respect to PBS to estimate the amount of

MNPs in suspension after 24 h. However, we can monitor the value of " at the peak-

frequency of the Brownian component of the MNPs in order to evaluate if nanoparticle-

protein interactions are restricting the rotation of the MNPs in solution during the

experimental time window of 24 h as shown in Figure 4-11.

Page 134: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

134

APS coated MNPs interact with blood and tumor stroma environments due their

positive charge and poor colloidal stability forming aggregates. Hence, rotation of

particles is restricted which results in the disappearance of the Brownian peak from the

frequency range used in this study. In contrast, for CMDx, capped CMDx and PEG

coated MNPs the Brownian peak does not disappear, suggesting that the physical

rotation of particles relaxing by the Brownian mechanism remains unaffected. This can

be explained by electrostatic repulsion between negatively charged blood proteins and

tumor stroma, and the negatively charged surface of the CMDx polymer shell. PEG

coated nanoparticles are neutral, hence do not interact with the charged biological

environment.

4.3.2 Visualization of Nanoparticle-Tumor Interactions Using Transmission Electron Microscopy

To characterize cellular distribution of the particles in tumor tissue ex-vivo, we

performed transmission electron microscopy after 30 min of particles incubation with

tumor tissue fragments. For these studies five MNPs from the DMS susceptibility

studies, namely positively charged PEI MNPs, commercial BNF-Starch 100 nm

(neutral), CMDx coated iron oxide MNPs (negative charged), capped CMDx coated iron

oxide MNPs (negative), and PEG coated iron oxide MNPs (neutral) were selected.

PEI MNPs showed a strong aggregation pattern in tissue after particle contact

ex-vivo (Figure 4-12). PEI coated MNPs were found in dark necrotic disintegrating cells.

These results agree with the DMS response studied in tumor tissue (no particle

rotation). There was no evidence of particle uptake in cancer cells or macrophages.

For the rest of the MNPs studied the distribution patterns in tumor tissue were

slightly different under TEM. Particularly, in the case of PEG coated particles (cIO-

Page 135: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

135

PEG), which were distributed into the tumor stromal tissue and cell bodies in a less

clustered form in comparison than the rest of the particles in this study.

We observed an agreement between TEM and DMS measurements in the case

of PEG coated MNPs (Figure 4-13) and BNF-starch MNPs (Figure 4-14) and which

showed endocytosis of MNPs after 30 min. It seems like BNF-Starch 100nn MNPs were

taken up by tumor associated macrophages and internalized in vesicles and lysosomal

compartments. CMDx and capped CMDx coated MNPs did not show agreement with

our previous DMS measurements (Figures 4-15 and 4-16 respectively). They look

aggregated but they are not restricted inside cell structures, which could be the reason

to observe particle mobility via DMS measurements after 24 h. These particles were

identified in several cancer cell structures and in dark necrotic disintegrating cells but

not in tumor associated macrophages.

4.4 Conclusions

The response of MNPs to an alternating magnetic field can be used to study

colloidal stability as a function of particle mobility in situ in blood and tumor tissue

environments via magnetic relaxation measurements. It was confirmed that the mobility

of particles in blood and particle distribution in tumor tissue depends strongly on

electrostatic interactions. First, it was identified that cationic particles aggregated in

tissue avoiding particle diffusion, which was observed as a reduction in the initial

susceptibility component as a function of time. Second, PEG coated nanoparticles as-

synthetized by the co-precipitation method remained stable up to 24 h observation in

blood and tissue with a size less than 100 nm. This agreed with previous biodistribution

results in our research group. Third, less interaction patterns were observed with CMDx

coated MNPs modified with sulfo-NHS-acetate in contrast with CMDx coated MNPs.

Page 136: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

136

Fourth, BNF-Starch MNPs aggregated in both blood and tumor tissue environments.

The results highlight that the dynamic response of MNPs in an alternating magnetic field

can be used to study colloidal stability as a function of particle mobility in situ in blood

and tumor tissue environments.

Page 137: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

137

Figure 4-1. Particle mobility assessment in blood and tumor tissue ex-vivo. 1) Cancer cell inoculation in a mice model, 2) tumor harvest and blood collection 30 days after cancer cells inoculation, 3) blood sample preparation, 4) tumor tissue sample preparation, and 5) dynamic magnetic susceptibility measurements at room temperature

Page 138: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

138

Figure 4-2. FTIR measurements for cIO-APS, cIO-CMDx, capped cIO-CMDx and cIOPEG magnetic nanoparticles

Page 139: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

139

Figure 4-3. Representative TEM micrographs and DMS spectra for PEI, CMDx, starch and PEG coated MNPs. TEM images show particle clusters and DMS measurements reveal that particles respond in the presence of an alternating magnetic field by physical rotation resulting in a Debye-like susceptibility plot

Page 140: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

140

Figure 4-4. Comparison between polymer coated particles selected in this study. a) ɀ-potential measurements and b) amine quantification results

Page 141: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

141

Figure 4-5. Equilibrium magnetization curves at 300 K for the PEI coated magnetic

nanoparticles (tICO-PEI) show a magnetization value about 125 A/m without mass normalization

Page 142: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

142

Figure 4-6. DMS spectra and lognormal diameter distributions for thermally blocked PEI,

CMDx and PEG coated MNPs in PBS, whole blood and tumor tissue

Page 143: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

143

Figure 4-7. Initial suceptibility values normalized with respect to samples in PBS and arithmetic mean hydrodynamic diameter – PEI, CMDx, PEG coated thermally blocked cobalt ferrite MNPs

Page 144: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

144

Figure 4-8. DMS spectra and lognormal diameter distributions for thermally blocked BNF-Starch 80 nm and 100 nm MNPs in PBS, whole blood and tumor tissue

Page 145: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

145

Figure 4-9. Normalized initial susceptibility values with respect to PBS – BNF-Starch thermally blocked iron oxide nanoparticles

Page 146: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

146

Figure 4-10. DMS spectra and lognormal diameter distributions for cIO-MNPs as-

synthetized by co-precipitation method with different surface functionalities – particles with mixed magnetic relaxation mechanisms

Page 147: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

147

Figure 4-11. Monitoring peak frequency displacement to the lower frequencies area upon contact with PBS, blood and

tumor tissue for MNPs with magnetic relaxation mixtures

Page 148: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

148

Figure 4-12. TEM micrographs of PEI coated cobalt ferrite nanoparticles in tumor tissue ex-vivo. Highly electron dense

black dots correspond to PEI coated nanoparticles. There were observed disintegrating cells (D), vesicles (V), and a swollen mitochondria (s). a) PEI coated MNPs adjacent to a cancer cell, b) Zoom in of an area full of MNPs (right) and delimitating disintegrating cells from the cancer cell cytoplasm (left)

Page 149: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

149

Figure 4-13. TEM micrograph of PEG coated iron oxide nanoparticles in tumor tissue ex-vivo. Highly electron dense black

dots correspond to PEG coated nanoparticles. a) Perinuclear membrane (Pm) with PEG coated MNPs, and a section with apoptotic cell bodies (A), b) Cancer cell showing part of the cell nucleus (N), mitochondria (Mi) and swollen mitochondria (S) and a lysosome (L) (left). Right side section shows a zoom in the lysosomal compartment with a highly electron dense area which may correspond to internalized PEG coated MNPs

Page 150: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

150

Figure 4-14. TEM micrographs of commercial BNF-Starch magnetic nanoparticles in

tumor tissue ex-vivo. Highly electron dense black quasi-squares correspond to BNF-Starch 100nm nanoparticles. a) Vesicle (V) with colocalized BNF starch MNPs (left), and zoom in of MNPs inside vesicles (right)

Page 151: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

151

Figure 4-15. TEM micrographs of CMDx coated iron oxide nanoparticles in tumor tissue ex-vivo. Highly electron dense

black quasi-spheres correspond to CMDx coated nanoparticles. a) Cancer cell (C) with some areas marked in squares showing CMDx coated MNPs, cell nucleolus (N) and lysosomal bodies (L) (left) , in the right side a zoom of the left top area is shown ,b) Stroma tissue (St) with CMDx coated MNPs

Page 152: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

152

Figure 4-16. TEM micrographs of Capped CMDx coated iron oxide nanoparticles in tumor tissue ex-vivo. Highly electron

dense black dots correspond to capped CMDx coated nanoparticles. a) Perinuclear membrane (Pm), a cancer cell, a lysosome (L) and Capped CMDx coated MNPs can be distinguished (left), b) zoom in of Capped CMDx coated MNPs

Page 153: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

153

CHAPTER 5 SUMMARY

This dissertation described the study of nanoparticle-protein interactions in situ

via dynamic magnetic susceptibility (DMS) measurements in optically transparent water-

glycerol mixtures at 37°C using SQUID-AC susceptometry with carboxymethyl dextran

(CMDx) coated cobalt ferrite MNPs. Due to the promising results using DMS

measurements in determining the colloidal stability of nanoparticles as a function of

nanoparticle-protein interactions, the use of this technique was extended to estimate the

rotational diffusion of highly concentrated protein solutions and particle mobility in whole

blood and tumor tissue using a commercial AC-susceptometer as follow:

5.1 Estimating Rotational Diffusivity of Magnetic Nanoparticles in Concentrated Protein Solutions

The rotational diffusivity of particle probes with reduced non-specific interactions

was monitored in concentrated protein solutions at different pH and salt conditions,

mimicking crowded biological environments. In this sense, particles were capable to

sense protein gelation at acidic conditions in the presence of sodium citrate ions.

5.2 Determining the Mobility of Polymer Coated Magnetic Nanoparticles in Blood and Tumor Tissue ex-vivo Using Magnetic Relaxation Measurements

Nanoparticle-blood interaction and nanoparticle-tumor stromal tissue interaction

studies were carried out with nanoparticles coated with different polymers to correlate

surface charge with the magnetic relaxation profile in the complex milieu, by detecting

changes in the hydrodynamic size of magnetic particles or aggregation patterns using

DMS measurements and transmission electron microscopy in tumor tissue fragments.

The polymer coatings evaluated represent commonly used polymers in biomedical

applications such as carboxymethyl dextran (negatively charged), branched

Page 154: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

154

polyethyleneimine (positively charged), hydroxyethyl starch and modified polyethylene

glycol (neutral polymers). Inorganic cores used in this study were thermally blocked

magnetic nanoparticles (cobalt ferrite core) and commercial thermally blocked iron oxide

nanoparticles. It was further demonstrated that this technique can be extended to

particle systems with a mixture of magnetization relaxation mechanisms.

5.3 Concluding Remarks

Nowadays, the design of new strategies to evaluate colloidal stability in complex

bio-environments have become valuable in targeted nanotherapeutics, since it has been

demonstrated that after systemic injection nanoparticle-blood interactions can

accelerate macrophage uptake with implications on solid tumor targeting and thus

limiting particle tumor accumulation. We expect that DMS measurements can be taken

into account more often for colloidal stability evaluations of polymer coated iron oxide

nanoparticles due to its reliability in situ by getting rid of artifacts associated to free

proteins at high protein at particle concentrations in contrast to the existent optical

techniques. This can take place without the limitation of optical transparency in sample

evaluation, and provides a useful assessment tool to design nanoparticles for

biomedical application with enhanced colloidally stability and minimal interactions with

blood, plasma proteins and tumor-stroma microenvironment.

Page 155: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

155

REFERENCES

[1] W. Seifriz, "An elastic value of protoplasm, with further observations on the viscosity of protoplasm," British Journal of Experimental Biology, vol. 2, pp. 1-11, Oct 1924.

[2] F. H. C. Crick and A. F. W. Hughes, "The physical properties of the cytoplasm - A study by means of the magnetic particle method. 1. Experimental," Experimental Cell Research, vol. 1, pp. 37-80, 1950.

[3] J. A. Nissim, "Intravenous administration of iron," Lancet, vol. 253, pp. 49-51, 1947.

[4] D. D. Stark, R. Weissleder, G. Elizondo, P. F. Hahn, S. Saini, L. E. Todd, et al., "Superparamagnetic iron oxide - Clinical applications as a contrast agent of MR imaging of the liver," Radiology, vol. 168, pp. 297-301, Aug 1988.

[5] R. K. Gilchrist, R. Medal, W. D. Shorey, R. C. Hanselman, J. C. Parrott, and C. B. Taylor, "Selective inductive heating of lymph nodes," Annals of Surgery, vol. 146, pp. 596-606, 1957 1957.

[6] MagForce. (2007-2015, 09/10/2016). Previous clinical treatment experience with NanoTherm™ therapy. Available: http://www.magforce.de/en/studien/uebersicht.html

[7] E. Amstad, M. Textor, and E. Reimhult, "Stabilization and functionalization of iron oxide nanoparticles for biomedical applications," Nanoscale, vol. 3, pp. 2819-43, Jul 2011.

[8] H. Arami, A. Khandhar, D. Liggitt, and K. M. Krishnan, "In vivo delivery, pharmacokinetics, biodistribution and toxicity of iron oxide nanoparticles," Chemical Society Reviews, vol. 44, pp. 8576-8607, 2015.

[9] W. R. Algar, D. E. Prasuhn, M. H. Stewart, T. L. Jennings, J. B. Blanco-Canosa, P. E. Dawson, et al., "The Controlled Display of Biomolecules on Nanoparticles: A Challenge Suited to Bioorthogonal Chemistry," Bioconjugate Chemistry, vol. 22, pp. 825-858, May 2011.

[10] Q. A. Pankhurst, J. Connolly, S. K. Jones, and J. Dobson, "Applications of magnetic nanoparticles in biomedicine," Journal of Physics D-Applied Physics, vol. 36, pp. R167-R181, Jul 2003.

[11] K. R. Hurley, H. L. Ring, H. Kang, N. D. Klein, and C. L. Haynes, "Characterization of Magnetic Nanoparticles in Biological Matrices," Analytical Chemistry, vol. 87, pp. 11611-11619, Dec 2015.

Page 156: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

156

[12] S. Sindhwani, A. M. Syed, S. Wilhelm, D. R. Glancy, Y. Y. Chen, M. Dobosz, et al., "Three-Dimensional Optical Mapping of Nanoparticle Distribution in Intact Tissues," Acs Nano, vol. 10, pp. 5468-5478, May 2016.

[13] S. Tenzer, D. Docter, J. Kuharev, A. Musyanovych, V. Fetz, R. Hecht, et al., "Rapid formation of plasma protein corona critically affects nanoparticle pathophysiology," Nat Nanotechnol, vol. 8, pp. 772-81, Oct 2013.

[14] A. Salvati, A. S. Pitek, M. P. Monopoli, K. Prapainop, F. B. Bombelli, D. R. Hristov, et al., "Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface," Nature Nanotechnology, vol. 8, pp. 137-143, Feb 2013.

[15] P. M. Kelly, C. Aberg, E. Polo, A. O'Connell, J. Cookman, J. Fallon, et al., "Mapping protein binding sites on the biomolecular corona of nanoparticles," Nat Nanotechnol, vol. 10, pp. 472-9, May 2015.

[16] M. Creixell, A. C. Bohorquez, M. Torres-Lugo, and C. Rinaldi, "EGFR-Targeted Magnetic Nanoparticle Heaters Kill Cancer Cells without a Perceptible Temperature Rise," Acs Nano, vol. 5, pp. 7124-7129, Sep 2011.

[17] B. Kozissnik, A. C. Bohorquez, J. Dobson, and C. Rinaldi, "Magnetic fluid hyperthermia: Advances, challenges, and opportunity," International Journal of Hyperthermia, vol. 29, pp. 706-714, 2013 2013.

[18] O. Veiseh, J. W. Gunn, and M. Zhang, "Design and fabrication of magnetic nanoparticles for targeted drug delivery and imaging," Adv Drug Deliv Rev, vol. 62, pp. 284-304, Mar 8 2010.

[19] J. Dobson, "Magnetic nanoparticles for drug delivery," Drug Development Research, vol. 67, pp. 55-60, Jan 2006.

[20] A. Riedinger, P. Guardia, A. Curcio, M. A. Garcia, R. Cingolani, L. Manna, et al., "Subnanometer Local Temperature Probing and Remotely Controlled Drug Release Based on Azo-Functionalized Iron Oxide Nanoparticles," Nano Letters, vol. 13, pp. 2399-2406, Jun 2013.

[21] M. H. Pablico-Lansigan, S. F. Situ, and A. C. S. Samia, "Magnetic particle imaging: advancements and perspectives for real-time in vivo monitoring and image-guided therapy," Nanoscale, vol. 5, pp. 4040-4055, 2013.

[22] H. W. Gu, P. L. Ho, K. W. T. Tsang, L. Wang, and B. Xu, "Using biofunctional magnetic nanoparticles to capture vancomycin-resistant enterococci and other gram-positive bacteria at ultralow concentration," Journal of the American Chemical Society, vol. 125, pp. 15702-15703, Dec 2003.

Page 157: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

157

[23] L. L. Bai, Y. M. Du, J. X. Peng, Y. Liu, Y. M. Wang, Y. L. Yang, et al., "Peptide-based isolation of circulating tumor cells by magnetic nanoparticles," Journal of Materials Chemistry B, vol. 2, pp. 4080-4088, 2014.

[24] E. G. Yarmola, Y. Shah, D. P. Arnold, J. Dobson, and K. D. Allen, "Magnetic Capture of a Molecular Biomarker from Synovial Fluid in a Rat Model of Knee Osteoarthritis," Annals of Biomedical Engineering, vol. 44, pp. 1159-1169, Apr 2016.

[25] S. Huth, J. Lausier, S. W. Gersting, C. Rudolph, C. Plank, U. Welsch, et al., "Insights into the mechanism of magnetofection using PEI-based magnetofectins for gene transfer," Journal of Gene Medicine, vol. 6, pp. 923-936, Aug 2004.

[26] J. Dobson, "Gene therapy progress and prospects: magnetic nanoparticle-based gene delivery," Gene Therapy, vol. 13, pp. 283-287, Feb 2006.

[27] M. Arsianti, M. Lim, C. P. Marquis, and R. Amal, "Assembly of Polyethylenimine-Based Magnetic Iron Oxide Vectors: Insights into Gene Delivery," Langmuir, vol. 26, pp. 7314-7326, May 2010.

[28] F. N. Al-Deen, J. Ho, C. Selomulya, C. Ma, and R. Coppel, "Superparamagnetic Nanoparticles for Effective Delivery of Malaria DNA Vaccine," Langmuir, vol. 27, pp. 3703-3712, Apr 2011.

[29] F. M. Kievit, O. Veiseh, C. Fang, N. Bhattarai, D. Lee, R. G. Ellenbogen, et al., "Chlorotoxin Labeled Magnetic Nanovectors for Targeted Gene Delivery to Glioma," Acs Nano, vol. 4, pp. 4587-4594, Aug 2010.

[30] K. Wang, F. M. Kievit, M. Jeon, J. R. Silber, R. G. Ellenbogen, and M. Q. Zhang, "Nanoparticle-Mediated Target Delivery of TRAIL as Gene Therapy for Glioblastoma," Advanced Healthcare Materials, vol. 4, pp. 2719-2726, Dec 2015.

[31] F. M. Kievit, O. Veiseh, N. Bhattarai, C. Fang, J. W. Gunn, D. Lee, et al., "PEI-PEG-Chitosan-Copolymer-Coated Iron Oxide Nanoparticles for Safe Gene Delivery: Synthesis, Complexation, and Transfection," Advanced Functional Materials, vol. 19, pp. 2244-2251, Jul 24 2009.

[32] D. P. Eisenberg, J. C. Bischof, and Y. Rabin, "Thermomechanical Stress in Cryopreservation Via Vitrification With Nanoparticle Heating as a Stress-Moderating Effect," Journal of Biomechanical Engineering-Transactions of the Asme, vol. 138, Jan 2016.

[33] J. Dobson, "Remote control of cellular behaviour with magnetic nanoparticles," Nature Nanotechnology, vol. 3, pp. 139-143, Mar 2008.

[34] H. Huang, S. Delikanli, H. Zeng, D. M. Ferkey, and A. Pralle, "Remote control of ion channels and neurons through magnetic-field heating of nanoparticles," Nat Nanotechnol, vol. 5, pp. 602-6, Aug 2010.

Page 158: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

158

[35] S. A. Stanley, J. E. Gagner, S. Damanpour, M. Yoshida, J. S. Dordick, and J. M. Friedman, "Radio-Wave Heating of Iron Oxide Nanoparticles Can Regulate Plasma Glucose in Mice," Science, vol. 336, pp. 604-608, May 2012.

[36] R. Chen, G. Romero, M. G. Christiansen, A. Mohr, and P. Anikeeva, "Wireless magnetothermal deep brain stimulation," Science, vol. 347, pp. 1477-1480, Mar 2015.

[37] D. Lavalette, M. A. Hink, M. Tourbez, C. Tetreau, and A. J. Visser, "Proteins as micro viscosimeters: Brownian motion revisited," Eur Biophys J, vol. 35, pp. 517-22, Aug 2006.

[38] A. D. Goncalves, C. Alexander, C. J. Roberts, S. G. Spain, S. Uddin, and S. Allen, "The effect of protein concentration on the viscosity of a recombinant albumin solution formulation," Rsc Advances, vol. 6, pp. 15143-15154, 2016.

[39] P. S. Sarangapani, S. D. Hudson, K. B. Migler, and J. A. Pathak, "The Limitations of an Exclusively Colloidal View of Protein Solution Hydrodynamics and Rheology," Biophysical Journal, vol. 105, pp. 2418-2426, Nov 2013.

[40] J. Jezek, M. Rides, B. Derham, J. Moore, E. Cerasoli, R. Simler, et al., "Viscosity of concentrated therapeutic protein compositions," Advanced Drug Delivery Reviews, vol. 63, pp. 1107-1117, Oct 2011.

[41] N. Inoue, E. Takai, T. Arakawa, and K. Shiraki, "Specific Decrease in Solution Viscosity of Antibodies by Arginine for Therapeutic Formulations," Molecular Pharmaceutics, vol. 11, pp. 1889-1896, Jun 2014.

[42] S. Amin, G. V. Barnett, J. A. Pathak, C. J. Roberts, and P. S. Sarangapani, "Protein aggregation, particle formation, characterization & rheology," Current Opinion in Colloid & Interface Science, vol. 19, pp. 438-449, Oct 2014.

[43] S. J. Wang, N. Zhang, T. Hu, W. G. Dai, X. Y. Feng, X. Y. Zhang, et al., "Viscosity-Lowering Effect of Amino Acids and Salts on Highly Concentrated Solutions of Two IgG1 Monoclonal Antibodies," Molecular Pharmaceutics, vol. 12, pp. 4478-4487, Dec 2015.

[44] P. Kheddo, M. J. Cliff, S. Uddin, C. F. van der Walle, and A. P. Golovanov, "Characterizing monoclonal antibody formulations in arginine glutamate solutions using H-1 NMR spectroscopy," Mabs, vol. 8, pp. 1245-1258, 2016.

[45] L. Nicoud, M. Lattuada, A. Yates, and M. Morbidelli, "Impact of aggregate formation on the viscosity of protein solutions," Soft Matter, vol. 11, pp. 5513-5522, 2015.

[46] M. M. Castellanos, J. A. Pathak, and R. H. Colby, "Both protein adsorption and aggregation contribute to shear yielding and viscosity increase in protein solutions," Soft Matter, vol. 10, pp. 122-131, 2014.

Page 159: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

159

[47] B. D. Connolly, C. Petry, S. Yadav, B. Demeule, N. Ciaccio, J. M. R. Moore, et al., "Weak Interactions Govern the Viscosity of Concentrated Antibody Solutions: High-Throughput Analysis Using the Diffusion Interaction Parameter," Biophysical Journal, vol. 103, pp. 69-78, Jul 2012.

[48] W. M. Saltzman, Drug Delivery Engineering Principles for Drug Therapy, 2001.

[49] B. Chertok, A. E. David, and V. C. Yang, "Brain tumor targeting of magnetic nanoparticles for potential drug delivery: Effect of administration route and magnetic field topography," Journal of Controlled Release, vol. 155, pp. 393-399, Nov 2011.

[50] L. Gutierrez, M. P. Morales, and F. J. Lazaro, "Prospects for magnetic nanoparticles in systemic administration: synthesis and quantitative detection," Physical Chemistry Chemical Physics, vol. 16, pp. 4456-4464, 2014.

[51] S. Chamorro, L. Gutierrez, M. P. Vaquero, D. Verdoy, G. Salas, Y. Luengo, et al., "Safety assessment of chronic oral exposure to iron oxide nanoparticles," Nanotechnology, vol. 26, May 2015.

[52] C. J. Cheng, G. T. Tietjen, J. K. Saucier-Sawyer, and W. M. Saltzman, "A holistic approach to targeting disease with polymeric nanoparticles," Nature Reviews Drug Discovery, vol. 14, pp. 239-247, Apr 2015.

[53] R. J. Ellis, "Macromolecular crowding: an important but neglected aspect of the intracellular environment," Current Opinion in Structural Biology, vol. 11, pp. 114–119, 2001.

[54] D. Rath and S. Panda, "Contribution of rotational diffusivity towards the transport of antigens in heterogeneous immunosensors," Analyst, vol. 140, pp. 6579-6587, 2015.

[55] G. A. Truskey, F. Yuan, and D. F. Katz, Transport Phenomena in Biological Systems, 2nd edition ed.: Pearson Education Inc., 2010.

[56] S. Gisladottir, T. Loftsson, and E. Stefansson, "Diffusion characteristics of vitreous humour and saline solution follow the Stokes Einstein equation," Graefes Archive for Clinical and Experimental Ophthalmology, vol. 247, pp. 1677-1684, Dec 2009.

[57] S. Ramanujan, A. Pluen, T. D. McKee, E. B. Brown, Y. Boucher, and R. K. Jain, "Diffusion and convection in collagen gels: Implications for transport in the tumor interstitium," Biophysical Journal, vol. 83, pp. 1650-1660, Sep 2002.

[58] A. Einstein, "The motion of elements suspended in static liquids as claimed in the molecular kinetic theory of heat," Annalen Der Physik, vol. 17, pp. 549-560, Jul 1905.

Page 160: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

160

[59] W. M. Deen, Analysis of Transport Phenomena, 2nd Edition ed., 2012.

[60] G. Gros, D. Lavalette, W. Moll, H. Gros, B. Amand, and F. Pochon, "Evidence for the rotational contribution to protein-facilitated proton transport," Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences, vol. 81, pp. 1710-1714, 1984.

[61] T. M. Squires and T. G. Mason, "Fluid Mechanics of Microrheology," Annual Review of Fluid Mechanics, vol. 42, pp. 413-438, 2010.

[62] S. H. Chen, F. Mallamace, C. Y. Mou, M. Broccio, C. Corsaro, A. Faraone, et al., "The violation of the Stokes-Einstein relation in supercooled water," Proceedings of the National Academy of Sciences of the United States of America, vol. 103, pp. 12974-12978, Aug 2006.

[63] A. Tuteja, M. E. Mackay, S. Narayanan, S. Asokan, and M. S. Wong, "Breakdown of the continuum Stokes-Einstein relation for nanoparticle diffusion," Nano Letters, vol. 7, pp. 1276-1281, May 2007.

[64] R. Poling-Skutvik, K. I. S. Mongcopa, A. Faraone, S. Narayanan, J. C. Conrad, and R. Krishnamoorti, "Structure and Dynamics of Interacting Nanoparticles in Semidilute Polymer Solutions," Macromolecules, vol. 49, pp. 6568-6577, Sep 2016.

[65] L. L. Josephson, E. M. Furst, and W. J. Galush, "Particle tracking microrheology of protein solutions," Journal of Rheology, vol. 60, pp. 531-540, Jul-Aug 2016.

[66] S. Wilhelm, A. J. Tavares, Q. Dai, S. Ohta, J. Audet, H. F. Dvorak, et al., "Analysis of nanoparticle delivery to tumours," Nature Reviews Materials, vol. 1, May 2016.

[67] H. Y. Zhou, J. L. Hao, S. Wang, Y. Zheng, and W. S. Zhang, "Nanoparticles in the ocular drug delivery," International Journal of Ophthalmology, vol. 6, pp. 390-396, Jun 2013.

[68] B. Shapiro, S. Kulkarni, A. Nacev, A. Sarwar, D. Preciado, and D. A. Depireux, "Shaping Magnetic Fields to Direct Therapy to Ears and Eyes," Annual Review of Biomedical Engineering, Vol 16, vol. 16, pp. 455-481, 2014.

[69] M. S. T. Qian Zhang, Anita Y. Carmichael-Baranauskas, Beth L. Caba, Michael A. Zalich, Yin-Nian Lin, O. Thompson Mefford, Richey M. Davis, and Judy S. Riffle, "Aqueous dispersions of magnetite nanoparticles complexed with copolyether dispersants: experiments and theory," Langmuir, vol. 23, pp. 6927-6936, 2007.

[70] C. Barrera, A. P. Herrera, and C. Rinaldi, "Colloidal dispersions of monodisperse magnetite nanoparticles modified with poly(ethylene glycol)," J Colloid Interface Sci, vol. 329, pp. 107-13, Jan 1 2009.

Page 161: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

161

[71] T. L. Moore, L. Rodriguez-Lorenzo, V. Hirsch, S. Balog, D. Urban, C. Jud, et al., "Nanoparticle colloidal stability in cell culture media and impact on cellular interactions," Chem Soc Rev, vol. 44, pp. 6287-305, Oct 7 2015.

[72] M. H. Kang and C. P. Reynolds, "Bcl-2 inhibitors: targeting mitochondrial apoptotic pathways in cancer therapy," Clin Cancer Res, vol. 15, pp. 1126-32, Feb 15 2009.

[73] M. Creixell, A. P. Herrera, M. Latorre-Esteves, V. Ayala, M. Torres-Lugo, and C. Rinaldi, "The effect of grafting method on the colloidal stability and in vitro cytotoxicity of carboxymethyl dextran coated magnetic nanoparticles," Journal of Materials Chemistry, vol. 20, pp. 8539-8547, 2010 2010.

[74] L. Santiago-Rodriguez, M. M. Lafontaine, C. Castro, J. Mendez-Vega, M. Latorre-Esteves, E. J. Juan, et al., "Synthesis, Stability, Cellular Uptake, and Blood Circulation Time of Carboxymethyl-Inulin Coated Magnetic Nanoparticles," J Mater Chem B Mater Biol Med, vol. 1, pp. 2807-2817, Jun 14 2013.

[75] S. Mondini, M. Leonzino, C. Drago, A. M. Ferretti, S. Usseglio, D. Maggioni, et al., "Zwitterion-Coated Iron Oxide Nanoparticles: Surface Chemistry and Intracellular Uptake by Hepatocarcinoma (HepG2) Cells," Langmuir, vol. 31, pp. 7381-90, Jul 7 2015.

[76] M. M. Yallapu, N. Chauhan, S. F. Othman, V. Khalilzad-Sharghi, M. C. Ebeling, S. Khan, et al., "Implications of protein corona on physico-chemical and biological properties of magnetic nanoparticles," Biomaterials, vol. 46, pp. 1-12, Apr 2015.

[77] L. Lartigue, C. Wilhelm, J. Servais, C. Factor, A. Dencausse, J.-C. Bacri, et al., "Nanomagnetic Sensing of Blood Plasma Protein Interactions with Iron Oxide Nanoparticles: Impact on Macrophage Uptake," Acs Nano, vol. 6, pp. 2665-2678, Mar 2012.

[78] A. C. Bohorquez and C. Rinaldi, "In Situ Evaluation of Nanoparticle-Protein Interactions by Dynamic Magnetic Susceptibility Measurements," Particle & Particle Systems Characterization, vol. 31, pp. 561-570, 2014.

[79] L. Gutierrez, S. Romero, G. B. da Silva, R. Costo, M. D. Vargas, C. M. Ronconi, et al., "Degradation of magnetic nanoparticles mimicking lysosomal conditions followed by AC susceptibility," Biomed Tech (Berl), vol. 60, pp. 417-25, Oct 2015.

[80] H. Arami, R. M. Ferguson, A. P. Khandhar, and K. M. Krishnan, "Size-dependent ferrohydrodynamic relaxometry of magnetic particle imaging tracers in different environments," Med Phys, vol. 40, p. 071904, Jul 2013.

[81] A. Jedlovszky-Hajdu, F. B. Bombelli, M. P. Monopoli, E. Tombacz, and K. A. Dawson, "Surface Coatings Shape the Protein Corona of SPIONs with Relevance to Their Application in Vivo," Langmuir, vol. 28, pp. 14983-14991, Oct 23 2012.

Page 162: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

162

[82] Y. Xu, Y. Qin, S. Palchoudhury, and Y. Bao, "Water-soluble iron oxide nanoparticles with high stability and selective surface functionality," Langmuir, vol. 27, pp. 8990-7, Jul 19 2011.

[83] W. T. Wang, X. Ji, H. B. Na, M. Safi, A. Smith, G. Palui, et al., "Design of a Multi-Dopamine-Modified Polymer Ligand Optimally Suited for Interfacing Magnetic Nanoparticles with Biological Systems," Langmuir, vol. 30, pp. 6197-6208, Jun 2014.

[84] K. Petersson, D. Ilver, C. Johansson, and A. Krozer, "Brownian motion of aggregating nanoparticles studied by photon correlation spectroscopy and measurements of dynamic magnetic properties," Analytica Chimica Acta, vol. 573, pp. 138-146, Jul 2006.

[85] L. Vroman and A. L. Adams, "Findings with recording ellipsometer suggesting rapid exchange of specific plasma proteins at the liquid/solid interfaces," Surface Science, vol. 16, pp. 438-&, 1969 1969.

[86] L. Vroman and A. L. Adams, "Identification of rapid changes at plasma-solid interfaces," Journal of biomedical materials research, vol. 3, pp. 43-67, 1969 1969.

[87] M. Jansch, P. Stumpf, C. Graf, E. Ruehl, and R. H. Mueller, "Adsorption kinetics of plasma proteins on ultrasmall superparamagnetic iron oxide (USPIO) nanoparticles," International Journal of Pharmaceutics, vol. 428, pp. 125-133, May 30 2012.

[88] P. Turbill, T. Beugeling, and A. A. Poot, "Proteins involved in the Vroman effect during exposure of human blood plasma to glass and polyethylene," Biomaterials, vol. 17, pp. 1279-1287, Jul 1996.

[89] J. Lazarovits, Y. Y. Chen, E. A. Sykes, and W. C. Chan, "Nanoparticle-blood interactions: the implications on solid tumour targeting," Chem Commun (Camb), vol. 51, pp. 2756-67, Feb 18 2015.

[90] M. A. Dobrovolskaia, P. Aggarwal, J. B. Hall, and S. E. McNeil, "Preclinical studies to understand nanoparticle interaction with the immune system and its potential effects on nanoparticle biodistribution," Molecular Pharmaceutics, vol. 5, pp. 487-495, Jul-Aug 2008.

[91] E. A. Vogler, "Protein adsorption in three dimensions," Biomaterials, vol. 33, pp. 1201-1237, Feb 2012.

[92] M. Mahmoudi, I. Lynch, M. R. Ejtehadi, M. P. Monopoli, F. B. Bombelli, and S. Laurent, "Protein-Nanoparticle Interactions: Opportunities and Challenges," Chemical Reviews, vol. 111, pp. 5610-5637, Sep 2011.

Page 163: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

163

[93] P. Aggarwal, J. B. Hall, C. B. McLeland, M. A. Dobrovolskaia, and S. E. McNeil, "Nanoparticle interaction with plasma proteins as it relates to particle biodistribution, biocompatibility and therapeutic efficacy," Advanced Drug Delivery Reviews, vol. 61, pp. 428-437, Jun 21 2009.

[94] F. Meder, T. Daberkow, L. Treccani, M. Wilhelm, M. Schowalter, A. Rosenauer, et al., "Protein adsorption on colloidal alumina particles functionalized with amino, carboxyl, sulfonate and phosphate groups," Acta Biomaterialia, vol. 8, pp. 1221-1229, Mar 2012.

[95] S. M. Griffiths, N. Singh, G. J. S. Jenkins, P. M. Williams, A. W. Orbaek, A. R. Barron, et al., "Dextran Coated Ultrafine Superparamagnetic Iron Oxide Nanoparticles: Compatibility with Common Fluorometric and Colorimetric Dyes," Analytical Chemistry, vol. 83, pp. 3778-3785, May 15 2011.

[96] M. P. Monopoli, C. Aberg, A. Salvati, and K. A. Dawson, "Biomolecular coronas provide the biological identity of nanosized materials," Nature Nanotechnology, vol. 7, pp. 779-786, Dec 2012.

[97] V. Filipe, A. Hawe, and W. Jiskoot, "Critical evaluation of Nanoparticle Tracking Analysis (NTA) by NanoSight for the measurement of nanoparticles and protein aggregates," Pharm Res, vol. 27, pp. 796-810, May 2010.

[98] C. Rocker, M. Potzl, F. Zhang, W. J. Parak, and G. U. Nienhaus, "A quantitative fluorescence study of protein monolayer formation on colloidal nanoparticles," Nat Nanotechnol, vol. 4, pp. 577-80, Sep 2009.

[99] S. Winzen, K. Koynov, K. Landfester, and K. Mohr, "Fluorescence labels may significantly affect the protein adsorption on hydrophilic nanomaterials," Colloids and Surfaces B: Biointerfaces, vol. 147, pp. 124-128, 2016.

[100] J. Romanowska, D. B. Kokh, and R. C. Wade, "When the Label Matters: Adsorption of Labeled and Unlabeled Proteins on Charged Surfaces," Nano Lett, vol. 15, pp. 7508-13, Nov 11 2015.

[101] L. Y. Chou and W. C. Chan, "Fluorescence-tagged gold nanoparticles for rapidly characterizing the size-dependent biodistribution in tumor models," Adv Healthc Mater, vol. 1, pp. 714-21, Nov 2012.

[102] L. Treuel, K. A. Eslahian, D. Docter, T. Lang, R. Zellner, K. Nienhaus, et al., "Physicochemical characterization of nanoparticles and their behavior in the biological environment," Phys Chem Chem Phys, vol. 16, pp. 15053-67, Aug 7 2014.

[103] Z. Krpeti, A. M. Davidson, M. Volk, R. Levy, M. Brust, and D. L. Cooper, "High-Resolution Sizing of Monolayer-Protected Gold Clusters by Differential Centrifugal Sedimentation " Acs Nano, vol. 7, pp. 8881–8890, 2013.

Page 164: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

164

[104] D. Walczyk, F. B. Bombelli, M. P. Monopoli, I. Lynch, and K. A. Dawson, "What the Cell "Sees" in Bionanoscience," Journal of the American Chemical Society, vol. 132, pp. 5761-5768, Apr 28 2010.

[105] W. Manchtle, "High-Resolution, Submicron Particle Size Distribution Analysis Using Gravitational-Sweep Sedimentation," Biophysical Journal, vol. 76, pp. 1080–1091, 1999.

[106] H. Amiri, L. Bordonali, A. Lascialfari, S. Wan, M. P. Monopoli, I. Lynch, et al., "Protein corona affects the relaxivity and MRI contrast efficiency of magnetic nanoparticles," Nanoscale, vol. 5, pp. 8656-65, Sep 21 2013.

[107] J. C. Bacri, R. Perzynski, D. Salin, and J. Servais, "Magnetic transient birefringence of ferrofluids : particle size determination," Journal de Physique, vol. 48, pp. 1385-1391, 1987.

[108] G. Meriguet, E. Dubois, V. Dupuis, and R. Perzynski, "Rotational arrest in a repulsive colloidal glass," Journal of Physics-Condensed Matter, vol. 18, pp. 10119-10132, Nov 2006.

[109] C. Riviere, C. Wilhelm, F. Cousin, V. Dupuis, F. Gazeau, and R. Perzynski, "Internal structure of magnetic endosomes," European Physical Journal E, vol. 22, pp. 1-10, Jan 2007.

[110] E. Hasmonay, A. Bee, J. C. Bacri, and R. Perzynski, "pH effect on an ionic ferrofluid: Evidence of a thixotropic magnetic phase," Journal of Physical Chemistry B, vol. 103, pp. 6421-6428, Aug 1999.

[111] C. Wilhelm, F. Gazeau, J. Roger, J. N. Pons, M. F. Salis, R. Perzynski, et al., "Binding of biological effectors on magnetic nanoparticles measured by a magnetically induced transient birefringence experiment," Physical Review E, vol. 65, p. 9, Mar 2002.

[112] K. Aurich, G. Glockl, S. Nagel, and W. Weitschies, "Magneto-Optical Relaxation Measurements of Functionalized Nanoparticles as a Novel Biosensor," Sensors, vol. 9, pp. 4022-4033, Jun 2009.

[113] F. J. Lazaro, A. R. Abadia, M. S. Romero, L. Gutierrez, J. Lazaro, and M. P. Morales, "Magnetic characterisation of rat muscle tissues after subcutaneous iron dextran injection," Biochim Biophys Acta, vol. 1740, pp. 434-45, Jun 10 2005.

[114] L. Gutiérrez, R. Mejías, D. F. Barber, S. Veintemillas-Verdaguer, C. J. Serna, F. J. Lázaro, et al., "Ac magnetic susceptibility study ofin vivonanoparticle biodistribution," Journal of Physics D: Applied Physics, vol. 44, p. 255002, 2011.

[115] J. Connolly and T. G. St Pierre, "Proposed biosensors based on time-dependent properties of magnetic fluids," Journal of Magnetism and Magnetic Materials, vol. 225, pp. 156-160, Apr 2001.

Page 165: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

165

[116] A. P. Astalan, F. Ahrentorp, C. Johansson, K. Larsson, and A. Krozer, "Biomolecular reactions studied using changes in Brownian rotation dynamics of magnetic particles," Biosensors & Bioelectronics, vol. 19, pp. 945-951, Mar 15 2004.

[117] S. H. Chung, A. Hoffmann, K. Guslienko, S. D. Bader, C. Liu, B. Kay, et al., "Biological sensing with magnetic nanoparticles using Brownian relaxation (invited)," Journal of Applied Physics, vol. 97, May 15 2005.

[118] A. P. Herrera, C. Barrera, Y. Zayas, and C. Rinaldi, "Monitoring colloidal stability of polymer-coated magnetic nanoparticles using AC susceptibility measurements," Journal of Colloid and Interface Science, vol. 342, pp. 540-549, Feb 15 2010.

[119] S. Sierra-Bermudez, L. P. Maldonado-Camargo, F. Orange, M. J. F. Guinel, and C. Rinaldi, "Assessing magnetic nanoparticle aggregation in polymer melts by dynamic magnetic susceptibility measurements," Journal of Magnetism and Magnetic Materials, vol. 378, pp. 64-72, Mar 2015.

[120] C. Barrera, V. Florian-Algarin, A. Acevedo, and C. Rinaldi, "Monitoring gelation using magnetic nanoparticles," Soft Matter, vol. 6, pp. 3662-3668, 2010 2010.

[121] V. L. Calero-DdelC, D. I. Santiago-Quinonez, and C. Rinaldi, "Quantitative nanoscale viscosity measurements using magnetic nanoparticles and SQUID AC susceptibility measurements," Soft Matter, vol. 7, pp. 4497-4503, 2011 2011.

[122] R. R. Merchant, L. Maldonado-Camargo, and C. Rinaldi, "In situ measurements of dispersed and continuous phase viscosities of emulsions using nanoparticles," Journal of Colloid and Interface Science, vol. 486, pp. 241-248, Jan 2017.

[123] L. Maldonado-Camargo and C. Rinaldi, "Breakdown of the Stokes-Einstein Relation for the Rotational Diffusivity of Polymer Grafted Nanoparticles in Polymer Melts," Nano Letters, vol. 16, pp. 6767-6773, Nov 2016.

[124] S. M. Dalibor Soukup, Eva Cespedes, Jon Dobson, and Neil D. Telling, "In Situ Measurement of Magnetization Relaxation of Internalized Nanoparticles in Live Cells," Acs Nano, vol. 9, pp. 231-249, 2015.

[125] R. Di Corato, A. Espinosa, L. Lartigue, M. Tharaud, S. Chat, T. Pellegrino, et al., "Magnetic hyperthermia efficiency in the cellular environment for different nanoparticle designs," Biomaterials, vol. 35, pp. 6400-11, Aug 2014.

[126] A. P. Khandhar, R. M. Ferguson, J. A. Simon, and K. M. Krishnan, "Enhancing cancer therapeutics using size-optimized magnetic fluid hyperthermia," J Appl Phys, vol. 111, pp. 7B306-7B3063, Apr 1 2012.

Page 166: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

166

[127] S. Gandhi, H. Arami, and K. M. Krishnan, "Detection of Cancer-Specific Proteases Using Magnetic Relaxation of Peptide-Conjugated Nanoparticles in Biological Environment," Nano Lett, vol. 16, pp. 3668-74, Jun 8 2016.

[128] R. L. D. Nikolaos Panagiotopoulos, Mandy Ahlborg, Gael Bringout, Christina Debbeler, Matthias Graeser, Christian Kaethner, Kerstin Lüdtke-Buzug, Hanne Medimagh, Jan Stelzner, Thorsten M Buzug, Jörg Barkhausen, Florian M Vogt and Julian Haegele, "Magnetic particle imaging: current developments and future directions," International Journal of Nanomedicine, vol. 10, pp. 3097-3114, 22 April 2015 2015.

[129] M. I. Shliomis, "Effective Viscosity of Magnetic Suspensions " SOVIET PHYSICS JETP vol. 34, pp. 1291-1295, December 1972 1972.

[130] M. I. Shliomis, "Ferrohydrodynamics: Testing a third magnetization equation," Physical Review E, vol. 64, Dec 2001.

[131] A. Chaves, F. Gutman, and C. Rinaldi, "Torque and Bulk Flow of Ferrofluid in an Annular Gap Subjected to a Rotating Magnetic Field," Journal of Fluids Engineering, vol. 129, p. 412, 2007.

[132] R. E. Rosensweig, Ferrohydrodynamics: Dover Publications Inc., 1985.

[133] L. Neel, "Influence des fluctuations thermiques sur laimantation de grains ferromagnetics tres fins," Comptes Rendus Hebdomadaires Des Seances De L Academie Des Sciences, vol. 228, pp. 664-666, 1949.

[134] R. E. Rosensweig, "Heating magnetic fluid with alternating magnetic field," Journal of Magnetism and Magnetic Materials, vol. 252, pp. 370-374, Nov 2002.

[135] P. Debye, Polar Molecules: Dover Publication, Inc., 1929.

[136] P. C. Fannin, B. K. P. Scaife, and S. W. Charles, "A study of the complex susceptibility of ferrofluids and rotational-motion," Journal of Magnetism and Magnetic Materials, vol. 65, pp. 279-281, Mar 1987.

[137] P. C. Fannin, S. W. Charles, and T. Relihan, "On the use of complex suceptibility data to complement magnetic viscosity measurements," Journal of Physics D-Applied Physics, vol. 27, pp. 189-193, Feb 1994.

[138] E. Roeben, L. Roeder, S. Teusch, M. Effertz, U. K. Deiters, and A. M. Schmidt, "Magnetic particle nanorheology," Colloid and Polymer Science, vol. 292, pp. 2013-2023, Aug 2014.

[139] D. E. Owens and N. A. Peppas, "Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles," International Journal of Pharmaceutics, vol. 307, pp. 93-102, Jan 2006.

Page 167: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

167

[140] M. Lundqvist, I. Sethson, and B. H. Jonsson, "Protein adsorption onto silica nanoparticles: Conformational changes depend on the particles' curvature and the protein stability," Langmuir, vol. 20, pp. 10639-10647, Nov 23 2004.

[141] E. Casals, T. Pfaller, A. Duschl, G. J. Oostingh, and V. Puntes, "Time Evolution of the Nanoparticle Protein Corona," Acs Nano, vol. 4, pp. 3623-3632, Jul 2010.

[142] A. Nimmagadda, K. Thurston, M. U. Nollert, and P. S. F. McFetridge, "Chemical modification of SWNT alters in vitro cell-SWNT interactions," Journal of Biomedical Materials Research Part A, vol. 76A, pp. 614-625, Mar 1 2006.

[143] P. P. Karmali and D. Simberg, "Interactions of nanoparticles with plasma proteins: implication on clearance and toxicity of drug delivery systems," Expert Opinion on Drug Delivery, vol. 8, pp. 343-357, Mar 2011.

[144] M. S. Ehrenberg, A. E. Friedman, J. N. Finkelstein, G. Oberdoerster, and J. L. McGrath, "The influence of protein adsorption on nanoparticle association with cultured endothelial cells," Biomaterials, vol. 30, pp. 603-610, Feb 2009.

[145] H. T. R. Wiogo, M. Lim, V. Bulmus, J. Yun, and R. Amal, "Stabilization of Magnetic Iron Oxide Nanoparticles in Biological Media by Fetal Bovine Serum (FBS)," Langmuir, vol. 27, pp. 843-850, Jan 18 2011.

[146] V. Ayala, A. P. Herrera, M. Latorre-Esteves, M. Torres-Lugo, and C. Rinaldi, "Effect of surface charge on the colloidal stability and in vitro uptake of carboxymethyl dextran-coated iron oxide nanoparticles," Journal of Nanoparticle Research vol. 15, p. 14, 2013.

[147] R. W. Eyler, E. D. Klug, and F. Diephuis, "Determination of Degree of Substitution of Sodium Carboxymethylcellulose," Analytical Chemistry, vol. 19, pp. 24-27, 1947 1947.

[148] Y. Cedeno-Mattei, O. Perales-Perez, M. S. Tomar, F. Roman, P. M. Voyles, and W. G. Stratton, "Tuning of magnetic properties in cobalt ferrite nanocrystals," Journal of Applied Physics, vol. 103, Apr 2008.

[149] J. d. A. Gomes, M. H. Sousa, F. A. Tourinho, R. Aquino, G. J. da Silva, J. Depeyrot, et al., "Synthesis of core-shell ferrite nanoparticles for ferrofluids: Chemical and magnetic analysis," Journal of Physical Chemistry C, vol. 112, pp. 6220-6227, Apr 24 2008.

[150] A. P. Herrera, C. Barrera, and C. Rinaldi, "Synthesis and functionalization of magnetite nanoparticles with aminopropylsilane and carboxymethyldextran," Journal of Materials Chemistry, vol. 18, pp. 3650-3654, 2008 2008.

[151] A. Lopez-Cruz, C. Barrera, V. L. Calero-DdelC, and C. Rinaldi, "Water dispersible iron oxide nanoparticles coated with covalently linked chitosan," Journal of Materials Chemistry, vol. 19, pp. 6870-6876, 2009 2009.

Page 168: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

168

[152] K. Gekko and S. N. Timasheff, "Mechanism of Protein Stabilization by Glycerol - Preferential Hydratation in Glycerol-Water Mixtures" Biochemistry, vol. 20, pp. 4667-4676, 1981 1981.

[153] V. Vagenende, M. G. S. Yap, and B. L. Trout, "Mechanisms of Protein Stabilization and Prevention of Protein Aggregation by Glycerol," Biochemistry, vol. 48, pp. 11084-11096, Nov 24 2009.

[154] W. C. Miles, J. D. Goff, P. P. Huffstetler, C. M. Reinholz, N. Pothayee, B. L. Caba, et al., "Synthesis and Colloidal Properties of Polyether-Magnetite Complexes in Water and Phosphate-Buffered Saline," Langmuir, vol. 25, pp. 803-813, Jan 20 2009.

[155] G. T. Hermanson, "Bioconjugate techniques," Bioconjugate techniques, pp. xxvii+785p-xxvii+785p, 1996 1996.

[156] V. Sharma, A. Jaishankar, Y. C. Wang, and G. H. McKinley, "Rheology of globular proteins: apparent yield stress, high shear rate viscosity and interfacial viscoelasticity of bovine serum albumin solutions," Soft Matter, vol. 7, pp. 5150-5160, 2011.

[157] L. Maldonado-Camargo, I. Torres-Díaz, A. Chiu-Lam, M. Hernández, and C. Rinaldi, "Estimating the contribution of Brownian and Néel relaxation in a magnetic fluid through dynamic magnetic susceptibility measurements," Journal of Magnetism and Magnetic Materials, vol. 412, pp. 223-233, 2016.

[158] J. Tafur, A. P. Herrera, C. Rinaldi, and E. J. Juan, "Development and validation of a 10 kHz-1 MHz magnetic susceptometer with constant excitation field," Journal of Applied Physics, vol. 111, Apr 1 2012.

[159] A. P. Astalan, C. Jonasson, K. Petersson, J. Blomgren, D. Ilver, A. Krozer, et al., "Magnetic response of thermally blocked magnetic nanoparticles in a pulsed magnetic field," Journal of Magnetism and Magnetic Materials, vol. 311, pp. 166-170, Apr 2007.

[160] J. W. Hubbard, F. Orange, M. J. F. Guinel, A. J. Guenthner, J. M. Mabry, C. M. Sahagun, et al., "Curing of a Bisphenol E Based Cyanate Ester Using Magnetic Nanoparticles as an Internal Heat Source through Induction Heating," Acs Applied Materials & Interfaces, vol. 5, pp. 11329-11335, Nov 2013.

[161] P. S. Sarangapani, S. D. Hudson, R. L. Jones, J. F. Douglas, and J. A. Pathak, "Critical Examination of the Colloidal Particle Model of Globular Proteins," Biophysical Journal, vol. 108, pp. 724-737, Feb 2015.

[162] A. Stradner, H. Sedgwick, F. Cardinaux, W. C. K. Poon, S. U. Egelhaaf, and P. Schurtenberger, "Equilibrium cluster formation in concentrated protein solutions and colloids," Nature, vol. 432, pp. 492-495, Nov 2004.

Page 169: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

169

[163] A. Grupi and A. P. Minton, "Concentration-Dependent Viscosity of Binary and Ternary Mixtures of Nonassociating Proteins: Measurement and Analysis," Journal of Physical Chemistry B, vol. 117, pp. 13861-13865, Nov 2013.

[164] C. Fernandez and A. P. Minton, "Effect of Nonadditive Repulsive Intermolecular Interactions on the Light Scattering of Concentrated Protein-Osmolyte Mixtures," Journal of Physical Chemistry B, vol. 115, pp. 1289-1293, Feb 2011.

[165] G. V. Barnett, W. Qi, S. Amin, E. N. Lewis, V. I. Razinkov, B. A. Kerwin, et al., "Structural Changes and Aggregation Mechanisms for Anti-Streptavidin IgG1 at Elevated Concentration," Journal of Physical Chemistry B, vol. 119, pp. 15150-15163, Dec 2015.

[166] G. V. Barnett, W. Qi, S. Amin, E. N. Lewis, and C. J. Roberts, "Aggregate structure, morphology and the effect of aggregation mechanisms on viscosity at elevated protein concentrations," Biophysical Chemistry, vol. 207, pp. 21-29, Dec 2015.

[167] C. Barrera, A. P. Herrera, N. Bezares, E. Fachini, R. Olayo-Valles, J. P. Hinestroza, et al., "Effect of poly(ethylene oxide)-silane graft molecular weight on the colloidal properties of iron oxide nanoparticles for biomedical applications," J Colloid Interface Sci, vol. 377, pp. 40-50, Jul 1 2012.

[168] D. Roberts, R. Keeling, M. Tracka, C. F. van der Walle, S. Uddin, J. Warwicker, et al., "Specific Ion and Buffer Effects on Protein-Protein Interactions of a Monoclonal Antibody," Molecular Pharmaceutics, vol. 12, pp. 179-193, Jan 2015.

[169] R. W. Chantrell, J. Popplewell, and S. W. Charles, "Measurements of particle-size distribution parameters in ferrofluids," Ieee Transactions on Magnetics, vol. 14, pp. 975-977, 1978.

[170] O. Esue, S. Kanai, J. Liu, T. W. Patapoff, and S. J. Shire, "Carboxylate-Dependent Gelation of a Monoclonal Antibody," Pharmaceutical Research, vol. 26, pp. 2478-2485, Nov 2009.

[171] O. Esue, A. X. Xie, T. J. Kamerzell, and T. W. Patapoff, "Thermodynamic and structural characterization of an antibody gel," Mabs, vol. 5, pp. 323-334, Mar-Apr 2013.

[172] J. Landers, L. Roeder, S. Salamon, A. M. Schmidt, and H. Wende, "Particle-Matrix Interaction in Cross-Linked PAAm-Hydrogels Analyzed by Mossbauer Spectroscopy," Journal of Physical Chemistry C, vol. 119, pp. 20642-20648, Sep 2015.

Page 170: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

170

[173] A. C. Alting, M. Weijers, E. H. A. De Hoog, A. M. van de Pijpekamp, M. A. C. Stuart, R. J. Hamer, et al., "Acid-induced cold gelation of globular proteins: Effects of protein aggregate characteristics and disulfide bonding on rheological properties," Journal of Agricultural and Food Chemistry, vol. 52, pp. 623-631, Feb 2004.

[174] L. R. S. Barbosa, M. G. Ortore, F. Spinozzi, P. Mariani, S. Bernstorff, and R. Itri, "The Importance of Protein-Protein Interactions on the pH-Induced Conformational Changes of Bovine Serum Albumin: A Small-Angle X-Ray Scattering Study," Biophysical Journal, vol. 98, pp. 147-157, Jan 2010.

[175] S. L. He, M. Huang, W. Ye, D. C. Chen, S. He, L. P. Ding, et al., "Conformational Change of Bovine Serum Albumin Molecules at Neutral pH in Ultra-Diluted Aqueous Solutions," Journal of Physical Chemistry B, vol. 118, pp. 12207-12214, Oct 2014.

[176] M. Dockal, D. C. Carter, and F. Ruker, "Conformational transitions of the three recombinant domains of human serum albumin depending on pH," Journal of Biological Chemistry, vol. 275, pp. 3042-3050, Feb 2000.

[177] K. Aoki and J. F. Foster, "Electrophoretic demonstration of the isomerization of bovine plasma albumin at low pH," Journal of the American Chemical Society, vol. 78, pp. 3538-3539, 1956.

[178] N. El Kadi, N. Taulier, J. Y. Le Huerou, M. Gindre, W. Urbach, I. Nwigwe, et al., "Unfolding and refolding of bovine serum albumin at acid pH: Ultrasound and structural studies," Biophysical Journal, vol. 91, pp. 3397-3404, Nov 2006.

[179] A. J. Minn, Y. B. Kang, I. Serganova, G. P. Gupta, D. D. Giri, M. Doubrovin, et al., "Distinct organ-specific metastatic potential of individual breast cancer cells and primary tumors," Journal of Clinical Investigation, vol. 115, pp. 44-55, Jan 2005.

[180] A. P. Herrera, L. Polo-Corrales, E. Chavez, J. Cabarcas-Bolivar, O. N. C. Uwakweh, and C. Rinaldi, "Influence of aging time of oleate precursor on the magnetic relaxation of cobalt ferrite nanoparticles synthesized by the thermal decomposition method," Journal of Magnetism and Magnetic Materials, vol. 328, pp. 41-52, 2013.

[181] M. A. Wang, M. L. Peng, W. Cheng, Y. L. Cui, and C. Chen, "A Novel Approach for Transferring Oleic Acid Capped Iron Oxide Nanoparticles to Water Phase," Journal of Nanoscience and Nanotechnology, vol. 11, pp. 3688-3691, Apr 2011.

[182] L. M.-C. Melissa Cruz-Acuña, Jon Dobson, Carlos Rinaldi, "From oleic acid-capped iron oxide nanoparticles to polyethyleneimine-coated single-particle magnetofectins," Journal of Nanoparticle Research, vol. 18, 2016.

Page 171: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

171

[183] R. Massart, "Preparation of aqueous magnetic liquids in alkaline and acid media," Ieee Transactions on Magnetics, vol. 17, pp. 1247-1248, 1981.

[184] M. Domenech, I. Marrero-Berrios, M. Torres-Lugo, and C. Rinaldi, "Lysosomal Membrane Permeabilization by Targeted Magnetic Nanoparticles in Alternating Magnetic Fields," Acs Nano, vol. 7, pp. 5091-5101, Jun 2013.

[185] E. A. Ellis, "Quetol 651: Not Just a Low Viscosity Resin," Microscopy Research and Technique, vol. 79, pp. 50-57, 2016.

[186] J. W. Wiseman, C. A. Goddard, D. McLelland, and W. H. Colledge, "A comparison of linear and branched polyethylenimine (PEI) with DCChol/DOPE liposomes for gene delivery to epithelial cells in vitro and in vivo," Gene Therapy, vol. 10, pp. 1654-1662, Sep 2003.

[187] J. W. Park, K. H. Bae, C. Kim, and T. G. Park, "Clustered Magnetite Nanocrystals Cross-Linked with PEI for Efficient siRNA Delivery," Biomacromolecules, vol. 12, pp. 457-465, Feb 2011.

[188] Y. Lee, S. H. Lee, J. S. Kim, A. Maruyama, X. Chen, and T. G. Park, "Controlled synthesis of PEI-coated gold nanoparticles using reductive catechol chemistry for siRNA delivery," J Control Release, vol. 155, pp. 3-10, Oct 10 2011.

[189] H. J. Lee, Y. T. C. Nguyen, M. Muthiah, H. Vu-Quang, R. Namgung, W. J. Kim, et al., "MR Traceable Delivery of p53 Tumor Suppressor Gene by PEI-Functionalized Superparamagnetic Iron Oxide Nanoparticles," Journal of Biomedical Nanotechnology, vol. 8, pp. 361-371, Jun 2012.

[190] M. Schwalbe, K. Pachmann, K. Hoffken, and J. H. Clement, "Improvement of the separation of tumour cells from peripheral blood cells using magnetic nanoparticles," Journal of Physics-Condensed Matter, vol. 18, pp. S2865-S2876, Sep 2006.

[191] W. Li, S. Tutton, A. T. Vu, L. Pierchala, B. S. Y. Li, J. M. Lewis, et al., "First-pass contrast-enhanced magnetic resonance angiography in humans using ferumoxytol, a novel ultrasmall superparamagnetic iron oxide (USPIO)-based blood pool agent," Journal of Magnetic Resonance Imaging, vol. 21, pp. 46-52, Jan 2005.

[192] H. L. Rodriguez-Luccioni, M. Latorre-Esteves, J. Mendez-Vega, O. Soto, A. R. Rodriguez, C. Rinaldi, et al., "Enhanced reduction in cell viability by hyperthermia induced by magnetic nanoparticles," International Journal of Nanomedicine, vol. 6, pp. 373-380, 2011 2011.

[193] M. P. Alvarez-Berríos, A. Castillo, J. Mendéz, O. Soto, C. Rinaldi, and M. Torres-Lugo, "Hyperthermic potentiation of cisplatin by magnetic nanoparticle heaters is correlated with an increase in cell membrane fluidity," IInternational Journal of Nanomedicine, vol. 11, pp. 1003-1013, 2013.

Page 172: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

172

[194] M. P. Alvarez-Berrios, A. Castillo, C. Rinaldi, and M. Torres-Lugo, "Magnetic fluid hyperthermia enhances cytotoxicity of bortezomib in sensitive and resistant cancer cell lines," International Journal of Nanomedicine, vol. 9, pp. 145-153, 2014 2014.

[195] F. Ahrentorp, A. Astalan, J. Blomgren, C. Jonasson, E. Wetterskog, P. Svedlindh, et al., "Effective particle magnetic moment of multi-core particles," Journal of Magnetism and Magnetic Materials, vol. 380, pp. 221-226, 2015.

[196] A. Jordan, R. Scholz, P. Wust, H. Schirra, T. Schiestel, H. Schmidt, et al., "Endocytosis of dextran and silan-coated magnetite nanoparticles and the effect of intracellular hyperthermia on human mammary carcinoma cells in vitro," Journal of Magnetism and Magnetic Materials, vol. 194, pp. 185-196, Apr 1999.

[197] X. Guo, P. Bandyopadhyay, B. Schilling, M. M. Young, N. Fujii, T. Aynechi, et al., "Partial acetylation of lysine residues improves intraprotein cross-linking," Analytical Chemistry, vol. 80, pp. 951-960, Feb 2008.

[198] E. K. U. Larsen, T. Nielsen, T. Wittenborn, L. M. Rydtoft, A. R. Lokanathan, L. Hansen, et al., "Accumulation of magnetic iron oxide nanoparticles coated with variably sized polyethylene glycol in murine tumors," Nanoscale, vol. 4, pp. 2352-2361, 2012 2012.

[199] C. Barrera, A. Herrera, Y. Zayas, and C. Rinaldi, "Surface modification of magnetite nanoparticles for biomedical applications," Journal of Magnetism and Magnetic Materials, vol. 321, pp. 1397-1399, 2009.

[200] J. P. Liu, Y. Z. Hsieh, D. Wiesler, and M. Novotny, "Design of 3-(4-carboxybenzoyl)-2-quinolinecarboxyaldehyde as a reagent for ultrasensitive determination of primary amines by capillary electrophoresis using laser fluorescence detection," Analytical Chemistry, vol. 63, pp. 408-412, Mar 1991.

[201] W. W. You, R. P. Haugland, D. K. Ryan, and N. P. Haugland, "3-(4-carboxybenzoyl)quinoline-2-carboxaldehyde, a reagent with broad dynamic range for the assay of proteins and lipoproteins in solution," Analytical Biochemistry, vol. 244, pp. 277-282, Jan 1997.

[202] N. Novatchev and U. Holzgrabe, "Evaluation of amino sugar, low molecular peptide and amino acid impurities of biotechnologically produced amino acids by means of CE," Journal of Pharmaceutical and Biomedical Analysis, vol. 28, pp. 475-486, May 2002.

[203] S. K. Samal, M. Dash, S. Van Vlierberghe, D. L. Kaplan, E. Chiellini, C. van Blitterswijk, et al., "Cationic polymers and their therapeutic potential," Chemical Society Reviews, vol. 41, pp. 7147-7194, 2012.

Page 173: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

173

[204] T. F. Martens, D. Vercauteren, K. Forier, H. Deschout, K. Remaut, R. Paesen, et al., "Measuring the intravitreal mobility of nanomedicines with single-particle tracking microscopy," Nanomedicine, vol. 8, pp. 1955-1968, Dec 2013.

Page 174: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

174

BIOGRAPHICAL SKETCH

Ana Carolina Bohorquez was born in Caracas, the capital city of Venezuela. She

pursued her Bachelor of Science in Chemical Engineering at the Central University of

Venezuela. As an undergraduate student, she received a scholarship from the Gran

Mariscal de Ayacucho foundation (Fundayacucho) to complete her undergraduate

studies. During her undergraduate studies, she did an internship in the Department of

New Business Development and Evaluation for the Orinoco Oil Belt at the Venezuelan

Petroleum Corporation (CVP), a division of Petroleos de Venezuela, S.A. (PDVSA)

which is the Venezuelan stated-owned oil and natural gas company. She was awarded

the Best Undergraduate Honor Thesis Distinction for her thesis project, “Feasibility

Study of Petroleum Coke Gasification Technologies for the Production of Electricity,

Hydrogen, and Liquid Hydrocarbons” in 2009. This distinction in her undergraduate

studies facilitated her to travel to San Salvador, El Salvador, where she was awarded

2nd place in the Oral Presentation Competition during the XV Latin American Congress

of Chemical Engineering Students 2009.

After acquiring some professional experience in Belcorp Cosmetics S.A as a

Warehouse Logistics Supervisor, she decided to pursue a Master of Science in

Chemical Engineering degree at the Universidad de Puerto Rico, Recinto Universitario

de Mayagüez in 2010. During her studies in Mayagüez, she conducted research

involving the synthesis of magnetic nanoparticles for biomedical applications under the

guidance of Dr. Magda Latorre-Esteves and Dr. Carlos Rinaldi. In fall 2012, she started

her doctoral degree in Biomedical Engineering at the University of Florida under the

supervision of Dr. Carlos Rinaldi with the aim to understand interactions of magnetic

materials with biological environments under the application of alternating magnetic

Page 175: © 2017 Ana Carolina Bohorquezufdcimages.uflib.ufl.edu/UF/E0/05/07/36/00001/BOHORQUEZ_A.pdf · Torres-Lugo for her technical advice and for sharing her expertise over the last few

175

fields. She received her Ph.D. from the University of Florida in the spring of 2017. Upon

graduating with her doctoral degree, Miss Bohorquez plans on working as a scientist in

the industry.