investigation of aminoglycoside induced nanoparticle self

62
University of Central Florida University of Central Florida STARS STARS Honors Undergraduate Theses UCF Theses and Dissertations 2018 Investigation of Aminoglycoside Induced Nanoparticle Self- Investigation of Aminoglycoside Induced Nanoparticle Self- Assemblies Assemblies Michael Leong University of Central Florida Part of the Biochemistry Commons, Biophysics Commons, and the Structural Biology Commons Find similar works at: https://stars.library.ucf.edu/honorstheses University of Central Florida Libraries http://library.ucf.edu This Open Access is brought to you for free and open access by the UCF Theses and Dissertations at STARS. It has been accepted for inclusion in Honors Undergraduate Theses by an authorized administrator of STARS. For more information, please contact [email protected]. Recommended Citation Recommended Citation Leong, Michael, "Investigation of Aminoglycoside Induced Nanoparticle Self-Assemblies" (2018). Honors Undergraduate Theses. 339. https://stars.library.ucf.edu/honorstheses/339

Upload: others

Post on 02-Jul-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Investigation of Aminoglycoside Induced Nanoparticle Self

University of Central Florida University of Central Florida

STARS STARS

Honors Undergraduate Theses UCF Theses and Dissertations

2018

Investigation of Aminoglycoside Induced Nanoparticle Self-Investigation of Aminoglycoside Induced Nanoparticle Self-

Assemblies Assemblies

Michael Leong University of Central Florida

Part of the Biochemistry Commons, Biophysics Commons, and the Structural Biology Commons

Find similar works at: https://stars.library.ucf.edu/honorstheses

University of Central Florida Libraries http://library.ucf.edu

This Open Access is brought to you for free and open access by the UCF Theses and Dissertations at STARS. It has

been accepted for inclusion in Honors Undergraduate Theses by an authorized administrator of STARS. For more

information, please contact [email protected].

Recommended Citation Recommended Citation Leong, Michael, "Investigation of Aminoglycoside Induced Nanoparticle Self-Assemblies" (2018). Honors Undergraduate Theses. 339. https://stars.library.ucf.edu/honorstheses/339

Page 2: Investigation of Aminoglycoside Induced Nanoparticle Self

INVESTIGATION OF AMINOGLYCOSIDE INDUCED NANOPARTICLE SELF-ASSEMBLIES

By

MICHAEL B. LEONG

A thesis submitted in partial fulfillment of the requirements for the Honors in the Major Program in Biomedical Sciences

in the College of Medicine and in the Burnett Honors College at the University of Central Florida

Orlando, Florida

Spring Term, 2018

Thesis Chair: William T. Self, Ph.D.

Page 3: Investigation of Aminoglycoside Induced Nanoparticle Self

ii

ABSTRACT

Aminoglycosides are a group of broad-spectrum antibiotics that, under neutral pH

conditions, carry a positive charge. The net cationic charge arises from the high number of

amino groups in the core structure of aminoglycosides. Previous studies performed have

shown that negatively charged citrate ligand-capped gold nanoparticles (AuNPs) can interact

with various biomolecules such as aminoglycosides. AuNPs bound to biomolecules have been

used in conjugation with various assaying techniques to detect and study compounds in vitro

and in vivo. AuNPs also have strong light scattering properties that can be used with a wide

variety of imaging and assaying techniques.

Our laboratory has previously performed experiments on the aminoglycoside antibiotic

ribostamycin sulfate. During this experiment, the concentration dependent rod-like assembly

of ribostamycin sulfate was characterized. This experiment used three analytical techniques in

conjunction with AuNPs: (1) dynamic light scattering (DLS), (2) UV-Vis absorption spectroscopy,

and (3) dark field optical microscope imaging (DFM). This suite of techniques was used to

analyze mixtures of ribostamycin sulfate at different concentration with different sized AuNPs.

The primary objective of this research was to determine if the techniques used to

characterize the self-assembly of ribostamycin sulfate could be generalized and applied to other

aminoglycoside antibiotics. The secondary objective of this research was to determine if other

aminoglycoside antibiotics formed rod-like assemblies.

Page 4: Investigation of Aminoglycoside Induced Nanoparticle Self

iii

This study demonstrated that AuNPs can be used to detect self-assembled oligomers for

different aminoglycoside antibiotics. In addition, this study also revealed that not all

aminoglycoside antibiotics will self assemble into rod-like oligomers similar to ribostamycin. It

was observed that the aminoglycoside antibiotic amikacin self assembled into rod-like

aggregates similar to ribostamycin sulfate but the aminoglycoside antibiotics neomycin sulfate

and streptomycin sulfate did not.

Page 5: Investigation of Aminoglycoside Induced Nanoparticle Self

iv

ACKNOWLEDGEMENT

I would like to express my gratitude to everyone who supported me throughout the course of my Honors in the Major thesis. I am thankful for their guidance, constructive criticism and

valuable advice during the project. I am sincerely grateful to them for sharing their expertise and knowledge.

UNIVERSITY OF CENTRAL FLORIDA

HONORS IN THE MAJOR COMMITTEE

Dr. Qun “Treen” Huo Associate Professor

Principal Investigator

Dr. William T. Self Associate Professor

Thesis Chair

Dr. Camilla Ambivero Assistant Professor

Committee Member

Special Thanks to Mr. Tianyu Zheng

Page 6: Investigation of Aminoglycoside Induced Nanoparticle Self

v

TABLE OF CONTENTS

LIST OF FIGURES ............................................................................................................................. vii

LIST OF ABBREVIATIONS ................................................................................................................. ix

INTRODUCTION ............................................................................................................................... 1

BACKGROUND ................................................................................................................................. 3

Aminoglycoside Antibiotics Overview ......................................................................................... 3

Aminoglycoside Uptake and Mechanism of Action .................................................................... 4

Aminoglycoside Toxicity .............................................................................................................. 7

Bacterial Resistance of Aminoglycosides .................................................................................... 9

Gold Nanoparticles .................................................................................................................... 10

Analytical Techniques Using AuNPs .......................................................................................... 12

Previous Research on Ribostamycin Sulfate ............................................................................. 14

METHODS ...................................................................................................................................... 18

Antibiotics and Chemicals ......................................................................................................... 18

Dynamic Light Scattering Assay................................................................................................. 18

UV-Vis Absorption Spectroscopy .............................................................................................. 19

Dark Field Optical Microscope Imaging .................................................................................... 20

Controls ..................................................................................................................................... 20

Page 7: Investigation of Aminoglycoside Induced Nanoparticle Self

vi

RESULTS AND DISCUSSION ........................................................................................................... 22

Dark Field Imaging ..................................................................................................................... 22

Amikacin in Nanopure Water .................................................................................................... 23

Streptomycin Sulfate in Nanopure Water ................................................................................. 27

Neomycin Sulfate in Nanopure Water ...................................................................................... 31

Neomycin Sulfate at Lower Concentrations in Nanopure Water ............................................. 34

Amikacin in 100 mM PB Buffer ................................................................................................. 38

Ribostamycin Sulfate in 100 mM PB Buffer .............................................................................. 42

CONCLUSION ................................................................................................................................. 46

LIST OF REFERENCES ..................................................................................................................... 48

Page 8: Investigation of Aminoglycoside Induced Nanoparticle Self

vii

LIST OF FIGURES

Figure 1. Chemical structure of the aminoglycosides analyzed in this study. ................................ 4

Figure 2. DLS results of previous ribostamycin sulfate study. ..................................................... 15

Figure 3. UV-Vis absorption spectroscopy of previous ribostamycin sulfate study. .................... 16

Figure 4. Dark field imaging results from previous ribostamycin study. ...................................... 17

Figure 5. Intensity-averaged size distribution curves of AuNP-amikacin mixtures in nanopure

water solvent ................................................................................................................................ 25

Figure 6. UV-Vis absorption spectra of AuNP-amikacin mixtures in nanopure water solvent .... 26

Figure 7. Intensity-averaged size distribution curves of AuNP-streptomycin sulfate mixtures in

nanopure water solvent ................................................................................................................ 29

Figure 8. UV-Vis absorption spectra of AuNP-streptomycin sulfate mixtures in nanopure water

solvent ........................................................................................................................................... 30

Figure 9. Intensity-averaged size distribution curves of AuNP-neomycin mixtures in nanopure

water solvent. ............................................................................................................................... 32

Figure 10. UV-Vis absorption spectra of AuNP-neomycin sulfate mixtures in nanopure water .. 33

Figure 11. Intensity-averaged size distribution curves of AuNP-neomycin sulfate mixtures in

nanopure water solvent. ............................................................................................................... 36

Figure 12. Intensity-averaged size distribution curves of AuNP-neomycin sulfate mixtures in

nanopure water solvent. ............................................................................................................... 37

Figure 13. Intensity-averaged size distribution curves of AuNP-amikacin mixtures in PB buffer

solvent. .......................................................................................................................................... 40

Page 9: Investigation of Aminoglycoside Induced Nanoparticle Self

viii

Figure 14. UV-Vis absorption spectra of AuNP-amikacin mixtures in PB buffer solvent. ............ 41

Figure 15. Intensity-averaged size distribution curves for AuNP-ribostamycin sulfate mixtures in

PB Buffer solvent........................................................................................................................... 44

Figure 16.UV-Vis absorption spectra of AuNP-ribostamycin sulfate mixtures in PB buffer solvent.

....................................................................................................................................................... 45

Page 10: Investigation of Aminoglycoside Induced Nanoparticle Self

ix

LIST OF ABBREVIATIONS

AuNP: Gold Nanoparticle

AAC: Aminoglycoside acetyltransferase

ANT: Aminoglycoside nucleotidyltransferase

APH: Aminoglycoside phosphotransferase

BCL-2: B-cell lymphoma-2

DFM: Dark Field Optical Microscope Imaging

DOS: Deoxystreptamine

DLS: Dynamic Light Scattering

EPR: Enhanced permeability and retention effect

ETC: Electron transport chain

EDPI: Energy dependent phase I

EDPII: Energy dependent phase II

JNK: c-jun N-terminal kinases

LSPR: Localized surface plasmon resonance

Ribo: Ribostamycin sulfate

ROS: Reactive oxygen species

Page 11: Investigation of Aminoglycoside Induced Nanoparticle Self

x

SPR: Surface plasmon resonance

ΔΨ: Membrane potential

Page 12: Investigation of Aminoglycoside Induced Nanoparticle Self

1

INTRODUCTION

Aminoglycosides are a group of antibiotics which are primarily used to treat Gram-

negative bacterial infections. For example, aminoglycosides are commonly used to treat a wide

range of bacterial respiratory infections such as Mycobacterium tuberculosis. Aminoglycosides

act by irreversibly binding the 16S subunit of the bacterial ribosome to inhibit protein

translation and by creating breaks within the outer membrane of the target bacteria to cause

the membrane to lyse.1 Despite the reversible nephrotoxic and permanent ototoxic side effects

of aminoglycosides, aminoglycosides have maintained widespread clinical use to this day. 2-3

However, an increasing number of bacterial strains have developed resistance to

aminoglycoside treatment such as methicillin resistant Staphylococcus aureus.4 Development

of aminoglycosides that can overcome these resistant strains requires an understanding of the

structure of aminoglycosides and the mechanism of action to inhibit protein synthesis.

AuNPs have a wide range of applications for biomedical research and therapeutics such

as compound analysis and drug delivery.5 The wide range of applications for AuNPs arises from

their strong light scattering properties that can be measured by different techniques.6 For

example, citrate ligand-capped AuNPs have been used to analyze aggregation of proteins with

dynamic light scattering assay, UV-Vis absorption spectroscopy and dark field optical

microscope imaging.7

Page 13: Investigation of Aminoglycoside Induced Nanoparticle Self

2

Our laboratory has previously characterized the concentration of dependent rod-like

assembly of ribostamycin sulfate using AuNPs. The characterization was observed by analyzing

ribostamycin-AuNP mixtures with DLS, UV-Vis absorption spectroscopy, and DFM.8 Three

antibiotic agents were analyzed in this study: amikacin, streptomycin sulfate, and neomycin

sulfate. Ribostamycin sulfate was analyzed in an additional experiment. Amikacin is a semi-

synthetic aminoglycoside derived from kanamycin and has a molecular weight of 585.603

g/mol. Streptomycin sulfate is an aminoglycoside sulfate salt that forms electrostatic

interactions with 1.5 sulfate ions per streptomycin and has a molecular weight of 728.69 g/mol.

Neomycin sulfate is an aminoglycoside sulfate salt that forms electrostatic interactions with

three sulfate ions per neomycin and has a molecular weight of 908.866 g/mol. Ribostamycin

sulfate is an aminoglycoside sulfate salt that forms electrostatic interactions with one sulfate

ion and has a molecular weight of 454.5 g/mol. All four aminoglycosides at neutral pH have a

positive charge which allows them to form a sulfate salt.

Page 14: Investigation of Aminoglycoside Induced Nanoparticle Self

3

BACKGROUND

Aminoglycoside Antibiotics Overview

In 1943, streptomycin was isolated from Streptomyces griseus. This was the first

aminoglycoside that had been isolated. Streptomycin was the first antibiotic to successfully

treat Mycobacterium tuberculosis, and it remains a common treatment for bacterial respiratory

infections.9 Other naturally occurring aminoglycosides have been isolated and include: (A)

neomycin was isolated from Streptomyces fradiae in 1949; (B) kanamycin was isolated from

Streptomyces kanamyceticus in 1957; (C) gentamicin was isolated from Micromonospora

purpurea in 1963; and (D) tobramycin was isolated from Streptomyces tenebrarius in 1967.10-13

To overcome bacterial resistance, previously discovered aminoglycosides have been modified

to yield semi-synthetic antibiotics such as amikacin, a kanamycin derivative.14

The structure of aminoglycosides contains aminated sugars that form glycosidic bonds

with a dibasic cyclitol. The core structure of an aminoglycoside is a neamine moiety which

contains Ring I that is connected to the fourth position of Ring II, the dibasic cyclitol. Other

sugars are attached to the core structure at position 5 or position 6 of Ring II. Common dibasic

cyclitols include streptidine, the dibasic cyclitol in streptomycin, and 2-deoxystreptamine (2-

DOS), the dibasic cyclitol for most aminoglycosides (Figure 1).

The two most common sub-classes of aminoglycosides are 4 and 5 disubstituted 2-DOS

(4,5-2-DOS), and 4 and 6 disubstituted 2-DOS (4,6-2-DOS) .15 An example of a 4,5-2-DOS

aminoglycoside is neomycin. Examples of 4,6-2-DOS aminoglycosides are amikacin and

Page 15: Investigation of Aminoglycoside Induced Nanoparticle Self

4

tobramycin.16 In 2-DOS aminoglycosides, the 2-DOS structure in combination with Ring I is

necessary to bind to the A site of the 16S rRNA subunit.17

Figure 1. Chemical structure of the aminoglycosides analyzed in this study.

Aminoglycoside Uptake and Mechanism of Action

Aminoglycosides are most effective in treating common strains of Gram-negative bacilli

and can also be used to treat certain gram-positive bacteria. Gram-negative bacilli that are

treated with aminoglycosides include Mycobacterium Tuberculosis and Pseudomonas

aeruginosa.1

Aminoglycoside uptake occurs in three phases: ionic binding, energy-dependent phase I

(EDPI), and energy-dependent phase II (EDPII).18

Neomycin Sulfate

Amikacin Ribostamycin Sulfate

Streptomycin Sulfate

Page 16: Investigation of Aminoglycoside Induced Nanoparticle Self

5

The ionic binding phase is unique as it is the only energy-independent phase of

aminoglycoside uptake. Aminoglycosides bind to anionic structures or sites on the surface of

the bacteria through reversible and concentration-dependent electrostatic interactions.

Aminoglycosides contain multiple positively charged amino groups on different rings allowing

aminoglycosides to electrostatically bind negatively charged structures located on the bacterial

envelope. Binding targets include lipopolysaccharides, phospholipids heads, teichoic acids, and

negatively charged membrane proteins.18 After binding, the aminoglycoside is transferred into

the cell in a process referred to as “self-promoted uptake.” During this process, Mg2+ bridges

are disrupted when aminoglycosides bind to Mg2+ binding sites on the bacterial cell membrane.

Aminoglycoside are able to enter the bacteria through these breaks in the membrane.19

The next phase of aminoglycoside uptake is EDPI. This phase is initiated after ionic

binding is completed. EDPI is the slow phase of aminoglycoside uptake and requires a

functioning electron transport chain (ETC) for the uptake of aminoglycosides. This phase of

aminoglycoside uptake requires a membrane potential (ΔΨ). It is currently unknown whether

the ΔΨ is generated directly by the ETC.20 It has been demonstrated in antibiotic resistant

bacteria that inducing a proton motive force will facilitate aminoglycoside uptake.21 For E. coli,

another factor that affects aminoglycoside uptake during EDPI is the growth rate of the E. coli

culture. A higher initial growth rate of E. coli has been shown to increase aminoglycoside

effectiveness. A higher initial growth rate will result in two main events. First, increased

growth rates will lead to higher ribosomal activity rates. Second, a higher initial growth rate

will cause an increase in aminoglycoside uptake with no change in aminoglycoside efflux. Both

events contribute to higher aminoglycoside accumulation within the bacteria. This results in a

Page 17: Investigation of Aminoglycoside Induced Nanoparticle Self

6

higher number of aminoglycoside compounds that can interact with a higher number of

ribosomes.22

The final phase of aminoglycoside uptake is EDPII. EDPII is the fast phase in

aminoglycoside uptake. During EDPII, aminoglycosides undergo linear transport across the

bacterial envelope. This process requires energy derived from the ETC, protein synthesis, and

ATP hydrolysis.23 Even though aminoglycosides ultimately inhibit protein synthesis, this final

phase requires aminoglycoside-sensitive ribosomes that are actively synthesizing proteins. If

active aminoglycoside-sensitive ribosomes are not present, EDPII will not occur.24

After aminoglycosides enter the bacteria, cell death is caused by the inhibition of

protein synthesis. Aminoglycosides accomplish this by targeting the ribosomes and by

compromising the integrity of the cytoplasmic membrane. After entering the bacteria,

aminoglycosides bind to the 16S rRNA subunit. The aminoglycoside can bind to the A site of the

polysome with differing conformations based on Ring II (dibasic cyclitol) interactions.

Aminoglycoside binding requires high affinity bonds with adenine 1492 (A1492) and guanine

1494 (G1494) of the 16S rRNA subunit.25 Aminoglycosides then form Watson-Crick base pairs

with the highly conserved adenine 1408 (A1408) and adenine 1493 (A1493). The Watson-Crick

base pairing creates a pocket in the polysome. The pocket causes the polysome to misread the

mRNA template and ultimately causes the termination of protein synthesis.26

Another possible pathway for aminoglycoside-induced protein synthesis inhibition may

be the stabilization of mRNA and tRNA in the A site. The tRNA will then be unable to move to

the P site and no other amino acids can be added to the polypeptide chain. This is caused by

changing codon-anticodon interactions.27 Although aminoglycosides primarily bind to the A

Page 18: Investigation of Aminoglycoside Induced Nanoparticle Self

7

site, aminoglycosides have been observed to bind to other parts of the polysome. For example,

it has been observed that neomycin and kanamycin form dimers which will bind to the

ribosome in two spots with high affinity.28

The bactericidal properties of aminoglycosides also are derived from the ability of

aminoglycosides to cause breaks in a bacteria’s cytoplasmic membrane. This occurs during the

ionic binding phase of aminoglycoside uptake. The “self-promoted uptake” increases

membrane permeability in a fashion similar to cationic detergents. The increased membrane

permeability ultimately allows more aminoglycosides to enter the cell. Compromising the

membrane allows free flow of smaller molecules which can disrupt the ion balance and

membrane potential. This results in the disruption of the bacterial cell’s metabolic processes.29

Aminoglycoside Toxicity

Aminoglycosides exhibit ototoxic and nephrotoxic side effects. Aminoglycoside-induced

damage to the inner ear is irreversible, but aminoglycoside-induced renal damage may be

reversed over time. The type and extent of damage depends on the aminoglycoside used.

Neomycin is one of the most toxic common aminoglycosides that can be administered to a

patient. Systematic administration of neomycin can potentially lead to permanent hearing

loss.1, 30

Nephrotoxicity is caused by aminoglycoside accumulation within the kidney’s epithelial

cells. This can result in impaired reabsorption of fluid, protein and solutes which may

ultimately result in nonoliguric renal failure. Upon administration, aminoglycosides will bind to

the brush border through electrostatic interactions with negatively charged phosphatidylserine.

Page 19: Investigation of Aminoglycoside Induced Nanoparticle Self

8

Next, megalin facilitates aminoglycoside uptake into the epithelial cells. Inside the epithelial

cell, aminoglycoside compounds start to localize in the Golgi apparatus, endosomes, and

lysosomes of epithelial cells in the proximal tubules.3 It has been observed that localization of

aminoglycosides within lysosomes causes an increase in the volume of lysosomes within hours

of aminoglycoside administration. Aminoglycosides inhibit enzymes such as lysosomal

phopsholipases and inhibit aggregation of phospholipids through electrostatic interactions.

These two factors are believed to be associated with necrosis of epithelial cells.31 The actual

mechanism for necrosis is unknown. It is believed that aminoglycoside-induced necrosis is a

major contributor to nephrotoxicity. In most cases, however, renal damage is reversible and

patients are able to recover.32

Ototoxicity is caused by aminoglycoside-induced damage to either the auditory or

vestibular organs. Aminoglycosides are categorized based on which part of the inner ear they

primarily affect. Streptomycin is mostly vestibulotoxic with some selective cochleotoxic effects.

Neomycin, kanamycin, and amikacin are mostly cochleotoxic.2 Aminoglycoside uptake into hair

cells is most likely mediated by ion channels.33 After uptake, aminoglycosides induce hair cells

to undergo apoptosis through the intrinsic apoptotic pathway. The entire pathway is unclear

but it has been observed that aminoglycosides increase the synthesis of reactive oxygen species

(ROS) which activate c-jun N-terminal kinases (JNK). JNK promotes pro-apoptotic B-cell

lymphoma-2 (BCL-2) proteins such as Bax and Bak. BCL-2 proteins will translocate into the

mitochondria and permeabilize the mitochondria resulting in the release of cytochrome c.

Cytochrome c initiates the caspase cascade that results in the death of the hair cell.2

Page 20: Investigation of Aminoglycoside Induced Nanoparticle Self

9

Bacterial Resistance of Aminoglycosides

Bacteria have several methods of resisting aminoglycoside activity. A common method

of resistance is to synthesize enzymes that modify the aminoglycoside’s structure. The three

classes of enzymes which affect all aminoglycosides are aminoglycoside acetyltransferases

(AACs), aminoglycoside nucleotidyltransferases (ANTs), and aminoglycoside

phosphotransferases (APHs).34 These enzymes recognize amino groups on an aminoglycoside

and modify the aminoglycoside using group transfers. Without the positively charged amino

groups, aminoglycosides are unable to properly bind to the 16S rRNA subunit. APHs add a

negatively charged phosphate group to carbon three which will form unfavorable interactions

with the negatively-charged phosphate backbone of the rRNA. ANTs and AACs catalyze group

transfers that prevent polysome binding due to steric hindrance.35

Another method of resistance is to decrease aminoglycoside uptake and develop a

mechanism for increased aminoglycoside efflux.34 By increasing the expression of anaerobic

respiratory pathway genes, the bacteria’s ATP yield decreases. This in turn reduces

aminoglycoside uptake during the energy dependent phases. Certain E. coli strains have

demonstrated active efflux of aminoglycosides in order to decrease accumulation.36-37

There are many different methods for aminoglycosides to subvert antibiotic resistance.

The most common method is the creation of semi-synthetic aminoglycosides such as amikacin.

Semi-synthetic aminoglycoside are able to avoid targeting by aminoglycoside modifying

enzymes.25 Another method of resistance is to increase the permeability of the bacterial

envelope. This can be accomplished by using other molecules that compromise bacterial

membrane integrity such as β-lactam antibiotics or plectasin.38

Page 21: Investigation of Aminoglycoside Induced Nanoparticle Self

10

Gold Nanoparticles

AuNPs are one of the most widely used metal nanoparticles in research. While AuNPs

have become more widely used within the past decade, knowledge of AuNPs and their

potential uses have been developed over the last 150 years. In 1857, Michael Faraday

synthesized the first colloidal AuNP solution by reducing gold chloride in the presence of a

phosphorus carbon disulfide solution.39 Currently, there are many different methods for

synthesizing AuNPs with varying sizes and shapes. The most common method is the Turkevich

method in which a gold salt undergoes citrate reduction to yield citrate-stabilized AuNPs.40

AuNPs can be used with various analytical techniques due to their optical properties.

The optical properties of AuNPs arise from their localized surface plasmon resonance (LSPR)

range. The LSPR of AuNPs is located within the visible spectrum of light (520-550 nm).

Exposing AuNPs to light from this range will result in extinction. The two main optical

properties of AuNPs are strong light scattering properties and strong light absorbance. AuNPs

demonstrate light scattering and absorption properties that are orders of magnitude higher

than conventional fluorescent and absorbing dyes. In addition, AuNPs exhibit light scattering

properties that are hundreds to thousands times stronger than polystyrene beads.6, 41

The strong light scattering exhibited by AuNPs is one of its most important properties.

AuNPs’ light scattering and absorption properties are dependent on the size, shape and

aggregation of the nanoparticle. AuNPs can vary in size from 2 nm to 100 nm. Commonly used

AuNP shapes are rods and spheres. AuNPs can be synthesized in other shapes such as cuboidal

or triangular plane AuNPs.42 Nanorods have demonstrated higher absorption and scattering

properties than spherical nanoparticles of equivalent size. Spherical AuNPs have demonstrated

Page 22: Investigation of Aminoglycoside Induced Nanoparticle Self

11

higher uptake and clearance from cells than nanorods.6, 43 These properties can be changed to

fit the needs of the assay.

AuNPs have a wide variety of applications for assaying biomolecules. AuNPs have been

utilized in protein studies, aggregation studies and colorimetric assays.44-46 Most applications

are linked to the ability of AuNPs to bind to a variety of compounds, their low reactivity, and

their various optical properties.47 An AuNP is commonly bound to a ligand to stabilize the

particle. This ligand determines the properties of the AuNP and what solvent it can be used in.

For example, a common ligand used is citrate or another molecule with a carboxylic acid

functional group that allows the AuNP to be water soluble.48 The unique surface properties of

AuNPs allow formation of ionic and hydrophobic interactions with various compounds. This

allows the AuNP to bind to numerous biomolecules such as antibodies, DNA, RNA and various

proteins.49

A major application of AuNPs is the delivery of molecules and compounds to targets

within the human body. AuNPs can be conjugated to small drug compounds which normally

have poor uptake to cells. The AuNPs aid in the uptake of the drug into a target cell by a

process known as “enhanced permeability and retention effect” (EPR effect).50 AuNPs can form

a monolayer around DNA and RNA to deliver the polynucleotide to target cells. AuNPs form

electrostatic interactions with the negatively charged DNA or RNA backbone and protect the

molecule from enzymes.48

AuNPs show promise for use in cancer therapeutics by aiding in drug delivery and

radiotherapy treatments.51 In addition, AuNPs are being used to develop new cancer

treatments such as siRNA and miRNA delivery to cut up mRNA for cancerous genes.52-53

Page 23: Investigation of Aminoglycoside Induced Nanoparticle Self

12

One currently debated topic is whether AuNPs are toxic to humans. Many studies

report that most AuNPs are non-toxic. However some studies have found that AuNPs of a

certain size combined with specific ligands can be toxic. The mechanism or reason for toxicity

remains unknown in many of these studies. While AuNPs appear to be usable in vivo, more

research needs to be done in this area.41, 54-55

Analytical Techniques Using AuNPs

The DLS assay utilizes the strong light scattering properties and Brownian motion of

AuNPs to detect the shape of the particle. AuNPs have light scattering properties that are

significantly higher than conventional tags such as fluorescent and absorbing dyes.6

The DLS assay shoots a laser through a sample and analyzes the scattered light. The

laser should be within the wavelength range that AuNPs exhibit LSPR. Once a laser hits the

sample, the light is scattered and picked up by a detector. The detector transforms the signal

and plots the intensity of the scattering as a function of time. This data is then analyzed by a

correlator which uses an autocorrelation function to determine various sized particles. Based

on the principles of Brownian motion, the larger and denser a particle is, the slower it will

move. Lighter and smaller particles will move through solution faster. After collecting the data

of the largest particles, the DLS software employs various fitting algorithms to determine the

size of the particle and graph this on an intensity distribution curve.

During this last step of the DLS assay, the hydrodynamic radius is calculated.56 The

hydrodynamic radius is the radius of a hard particle which has the same diffusion rate. This is

recorded by the DLS assay as the Z average diameter. The hydrodynamic radius is based on the

Page 24: Investigation of Aminoglycoside Induced Nanoparticle Self

13

dominating or largest light scattering source in the solution. The Stokes-Einstein equation is

used to calculate the hydrodynamic radius 𝑅𝐻. The variables that can be manipulated are

Boltzmann’s constant 𝑘, absolute temperature 𝑇, dynamic viscosity ƞ, pi 𝜋, and the diffusion

coefficient 𝐷. For our experiment, all variables are kept constant except the diffusion

coefficient. This allows manipulation of just the diffusion coefficient in order to yield the

hydrodynamic radius.57 The optical properties of an AuNP change when the AuNP interacts

with another molecule. This change in the optical properties results in a change in the

hydrodynamic radius58

Equation 1. Stokes-Einstein equation.

The presence of AuNPs and AuNP-aminoglycoside interactions affects the hydrodynamic

diameter because pure aminoglycosides have little to no light scattering properties. DLS

analysis of pure aminoglycoside dissolved in nanopore water, at experimental concentrations

and up to 50 mM, gave photon count rates barely above background noise. UV-Vis absorption

spectroscopy analysis of pure aminoglycosides dissolved in nanopore water failed to detect the

contents of the solution. Both observations demonstrate that any recorded measurements

were due to AuNPs or AuNP-aminoglycoside interactions.

DLS assay data is displayed in a distribution curve. One peak is characteristic of a

spherical molecule. The Z average diameter represents the average particle size. Two peaks

are characteristic of rod-like aggregation. The smaller peak located at a smaller Z average

𝑅𝐻 =𝑘𝑇

6𝜋ƞ𝐷

Page 25: Investigation of Aminoglycoside Induced Nanoparticle Self

14

diameter on the intensity distribution curve is a result of the rotational diffusion of the particle.

The larger peak located at a larger Z average diameter on the intensity distribution curve is a

result of the translational diffusion of the particle.59-60

UV-Vis absorption spectroscopy measures the absorption spectra of the AuNP or the

AuNP-aminoglycoside sample. UV-Vis absorption data is displayed in an absorption spectra.

One surface plasmon resonance (SPR) peak is characteristic of a sphere or spherical particle.

Rod-like particles or aggregates will display two peaks. There will be one peak within the SPR

range of the AuNP and a second, smaller, peak at a longer wavelength.60

Dark field optical microscope imaging uses an optical microscope combined with a dark

field imaging accessory to visualize AuNPs. The strong light scattering properties of the AuNPs

combined with their resistance to photobleaching allows visualization with DFM. Capturing the

AuNP-aminoglycoside mixture solution on the microscope slide allows the viewing of AuNPs

suspended in solution. This allows the viewer to see the Brownian motion of the AuNPs.61

Previous Research on Ribostamycin Sulfate

Our laboratory previously performed research on ribostamycin sulfate. The goal of this

study was to analyze ribostamycin sulfate using AuNPs. During this experiment it was observed

that ribostamycin sulfate self assembled into rod-like aggregates in a concentration dependent

manner. DLS analysis of ribostamycin sulfate with 40 nm and 60 nm AuNPs observed rod-like

aggregation at 1 mM and 500 µM concentrations. All other ribostamycin-AuNP mixtures did

not exhibit any peak shifts. UV-Vis absorption spectroscopy analysis revealed that 1 mM and

500 µM concentrations ribostamycin sulfate in the presence of 40 nm and 60 nm AuNPs

Page 26: Investigation of Aminoglycoside Induced Nanoparticle Self

15

displayed two peaks. All other mixtures displayed peaks resembling the control. The two peaks

showed that ribostamycin sulfate self-assembled into rod-like aggregates. These results were

confirmed by DFM. DFM pictures of ribostamycin sulfate showed rod-like aggregation. In

addition, even samples that did not have any detectable aggregation showed some kind of

aggregate formation. These structures did not have any definite shape.

Figure 2. DLS results of previous ribostamycin sulfate study.

Page 27: Investigation of Aminoglycoside Induced Nanoparticle Self

16

Figure 3. UV-Vis absorption spectroscopy of previous ribostamycin sulfate study.

Page 28: Investigation of Aminoglycoside Induced Nanoparticle Self

17

Figure 4. Dark field imaging results from previous ribostamycin study.

AuNP-40 nm AuNP-60 nm AuNP-80 nm AuNP-100 nm

Page 29: Investigation of Aminoglycoside Induced Nanoparticle Self

18

METHODS

Antibiotics and Chemicals

The aminoglycosides used in this experiment were neomycin sulfate, amikacin,

streptomycin sulfate and ribostamycin sulfate. The size and concentration of the following

nanoparticle suspensions were used: 40 nm (9.00 x 10 10 particle /mL), 60 nm (2.60 x 10 10

particle/mL), 80 nm (1.10 x 10 10 particle/mL) and 100 nm (5.60 x 10 10 particle/mL).

The four different citrate ligand-capped AuNP suspensions were purchased from Ted

Pella Inc. and were manufactured by British Biocell International The antibiotics used in this

experiment were purchased from MP Biomedicals. The antibiotics were stored in solid powder

form in a refrigerator at 4 °C. The PB buffer was diluted down to a 100 µM with nanopure

water from a 1 M stock and was stored in a refrigerator at 4 °C.

Dynamic Light Scattering Assay.

The analysis of each AuNP-aminoglycoside mixture solution was initiated with the DLS

assay. The DLS Assay was performed with a Zetasizer Nano ZS90 DLS system with a green laser

(532 nm, 4 mW) and an avalanche photodiode detector (quantum efficiency >50% at 532 nm).

The Malvern DTS 5.10 software and 7.10 software were utilized for data processing and

analysis. The detector was set at a 90 ° angle.

The scattering light intensity was measured and quantified as photons counted per

second. This was displayed by the software in kilo counts per second (kcps). The incident

power of the laser was changed based on the size of the nanoparticle so that the scattering

Page 30: Investigation of Aminoglycoside Induced Nanoparticle Self

19

light intensity readings were maintained between 100 kcps and 1000 kcps. The laser was

adjusted by an attenuator that is built into the Zetasizer. The level of attenuation was displayed

as an attenuation number and each number corresponded with a specific laser power. The

attenuation numbers used in this experiment were 10 (1.2 mW), 9 (0.4 mW), 8 (0.12 mW), and

7 (0.04 mW). Attenuation numbers 10 and 9 were used when analyzing antibiotic-AuNP

mixtures that contained 40 nm AuNPs and 60 nm AuNPs, respectively. Attenuation numbers 8

and 7 were used when analyzing antibiotic-AuNP mixtures that contained 80 nm AuNPs and

100 nm AuNPs, respectively. When taking a sample measurement, two separate

measurements were taken with a 20 second fixed run time.

Each aminoglycoside was dissolved in either nanopure water or PB buffer. The

dissolved antibiotic was then diluted to 1 mM, 500 µM, 250 µM, 100 µM, 50 µM, 20 µM, or 10

µM concentrations. For each experiment, 6 µL of the aminoglycoside dilution was mixed with

100 µL of a AuNP suspension in an Eppendorf microtube. This mixture solution was then

incubated for ten minutes. Shortly before the ten minute mark, the mixture solution was

transferred to a cuvette (catalog number 67.758, Sarstedt, Germany). Immediately after 10

minutes of incubation had transpired, four DLS measurements were taken at 2 minute intervals.

This was repeated for each sample.

UV-Vis Absorption Spectroscopy

The second step in the analysis of each AuNP-aminoglycoside mixture solution utilized

UV-Vis absorption spectroscopy. The UV-Vis spectroscopy analysis was performed with a

NanoDrop 2000 UV-Vis spectrophotometer purchased from Thermo Scientific. For each

Page 31: Investigation of Aminoglycoside Induced Nanoparticle Self

20

experiment, 2 µL of each AuNP-aminoglycoside mixture solution was pipetted onto the

pedestal. Two measurements were recorded for each sample and the average of those

measurements was recorded for results.

Dark Field Optical Microscope Imaging

The final step for analysis of each AuNP-aminoglycoside mixture solution utilized dark

field optical microscope imaging (DFM). DFM was performed using an Olympus BX51

microscope with a dark field condenser, oil dispersed objective lens (100x/1.3) and a Soft

Imaging System’s Colorview III camera with AnalySIS Life Sciences software. Slides were

prepared by pipetting 8 µL of the aminoglycoside-AuNP mixture solution onto a glass

microscope slide. The sample was covered with a glass cover, sealed, and mounted on the

stage. An image was then taken with the camera and exported to a computer.

Controls

Each of the four aminoglycoside antibiotics (ribostamycin sulfate, amikacin, neomycin

sulfate and streptomycin sulfate) were analyzed against a pure AuNP solution that acted as a

control. Each control presented a single peak indicating that each control was comprised of

only spherical nanoparticles. The AuNP control solutions were prepared for nanoparticles with

average Z diameters of 40 nm, 60 nm, 80 nm or 100 nm. These controls were performed for

every tested set of AuNP-aminoglycoside mixture solutions (Figures 5, 7, 9, 11, 13, 15).

It was observed that the UV-Vis absorption spectra of the pure AuNP controls exhibited

a single SPR peak of approximately 530 nm for 40 nm AuNPs, 540 nm for 60 nm AuNPs, 550 nm

Page 32: Investigation of Aminoglycoside Induced Nanoparticle Self

21

for 80 nm AuNPs and 580 nm for 100 nm AuNPs. These controls were performed for every

tested set of AuNP-aminoglycoside mixture solutions (Figures 6, 8, 10, 12, 14, 16). It should be

noted that after the amikacin 80 nm trial, the 80 nm control sometime presented two peaks.

This anomaly may have been caused by contamination of the 80 nm GNP solution. During the

amikacin-100 nm GNP trials, there was some contamination of the 100 nm GNP solution but

this was corrected in later trials. Additional trials were not repeated to account for the

anomalies due to time restrictions.

Page 33: Investigation of Aminoglycoside Induced Nanoparticle Self

22

RESULTS AND DISCUSSION

The hypothesis was AuNPs could be used to analyze aminoglycosides in order to

characterize rod -like aggregation in a concentration dependent manner. It was proposed that

citrate ligand-capped AuNPs could form electrostatic charges with any aminoglycoside because

the positively charged amino groups would be attracted to the negatively charged AuNPs. The

formation of a complex between the aminoglycosides and AuNPs would allow the analysis of

the aminoglycoside. This hypothesis was derived from previous research performed by our

laboratory though the analysis of ribostamycin sulfate using DLS assay, UV-Vis absorption

spectroscopy and dark field optical microscope imaging. To test this hypothesis, the same

protocol used to analyze ribostamycin sulfate was performed on several other aminoglycosides.

Dark Field Imaging

DFM images were recorded for amikacin in order to confirm rod-like aggregates. The

DFM confirmed the rod-like behavior of amikacin. The images were similar to those of

ribostamycin sulfate (Figure 4). Unfortunately due to university computer issues, the DFM

images that had been taken were lost. Additional trials to collect these images were not

performed due to time restrictions.

Page 34: Investigation of Aminoglycoside Induced Nanoparticle Self

23

Amikacin in Nanopure Water

Figure 5 contains the DLS data that is displayed as a intensity-averaged size distribution

curve. When amikacin is mixed with 40 nm and 60 nm AuNP solutions, two strong peaks were

observed (Figure 5-A, B). The 40 AuNP nm mixture displayed a second peak around the 8 nm

mark and the 60 nm AuNP mixture displayed a second peak that ranged from the 7 nm to 12

nm mark. The 80 nm and 100 nm AuNP solutions mixed with amikacin sometimes displayed a

smaller second peak but this may have been caused by contamination and not by amikacin-

AuNP interactions. Amikacin mixed with 80 nm and 100 nm AuNP solutions each exhibited one

strong peak which was almost identical to the peaks exhibited by the pure AuNP solution

(Figure 5-C, D). This may have been caused by the larger sized AuNPs mixed with amikacin not

being able to form significant sized anisotropic aggregates.

The UV-Vis absorption spectra data followed the same trend as the DLS data. It was

observed that 40 nm and 60 nm AuNPs mixed with amikacin generated two peaks (Figure 6-A,

B). The larger of the two peaks presented itself at around 540 nm. The second peak was

observed at longer wavelengths that ranged between 600 nm to 700 nm. Similar to the DLS

assay results, amikacin mixed with the 80 nm and 100 nm AuNPs presented one peak that was

similar to the peak displayed by the pure AuNP solution (Figure 6-C, D).

DFM provided further evidence of the anisotropic assembly of amikacin. There were

multiple self-assembled linear anisotropic particles. It was observed that 40 nm AuNP-amikacin

and 60 nm AuNP-amikacin mixtures contained visible aggregates at all concentrations. The

Page 35: Investigation of Aminoglycoside Induced Nanoparticle Self

24

AuNP-amikacin complexes displayed stronger light scattering than the pure AuNP solutions

under dark field imaging (see Figure 4 for reference).

The results of the DLS assay, UV-Vis absorption spectroscopy and DFM suggest that

amikacin displays concentration-dependent linear self assembly. Amikacin behaved similar to

ribostamycin sulfate. Our observations further support the proposal that amikacin can self

assemble as a linear rod-like aggregate.

There were some observed differences between the behavior of amikacin and

ribostamycin sulfate. Unlike amikacin, only 1 mM or 500 µM concentrations of ribostamycin

sulfate solution mixed with 40 nm or 60 nm AuNPs presented two peaks when analyzed with

the DLS assay and UV-Vis spectroscopy. In contrast to ribostamycin sulfate, amikacin displayed

two peaks at all tested concentrations when mixed with 40 nm and 60 nm AuNPs. This

observation suggests that amikacin can form aggregates at significantly lower concentrations

than ribostamycin sulfate. One possible explanation is that amikacin has one more amino

group than ribostamycin sulfate (Figure 1). This would allow amikacin to form more

electrostatic interactions with other molecules.

Page 36: Investigation of Aminoglycoside Induced Nanoparticle Self

25

Figure 5. Intensity-averaged size distribution curves of AuNP-amikacin mixtures in nanopure water solvent. The 40 nm (A) and 60 nm (B) AuNP mixtures with amikacin exhibited two peaks at all tested concentrations indicating rod-like self assembly. The 80 (C) nm and 100 (D) nm AuNP mixtures with amikacin did not exhibit any peak shifts.

Page 37: Investigation of Aminoglycoside Induced Nanoparticle Self

26

Figure 6. UV-Vis absorption spectra of AuNP-amikacin mixtures in nanopure water solvent. This supported the DLS measurements. The 40 nm (A) and 60 nm (B) AuNP mixtures displayed two peaks. The 80 nm (C) and 100 nm (D) AuNP mixtures displayed peaks similar to the control.

Page 38: Investigation of Aminoglycoside Induced Nanoparticle Self

27

Streptomycin Sulfate in Nanopure Water

Figure 7 contains the intensity-averaged size distribution curves of streptomycin sulfate

mixture solutions. AuNP-streptomycin sulfate mixtures consistently presented one strong peak

during DLS assay analysis. It was often observed that streptomycin sulfate presented peaks

with a Z average diameter much larger than pure AuNP. It was observed that 40 nm AuNP-

streptomycin sulfate mixtures exhibited peaks at around 710 nm (Figure 7-A), 60 nm AuNP-

streptomycin sulfate mixtures exhibited peaks at around 530 nm (Figure 7-B), 80 nm AuNP-

streptomycin sulfate mixtures exhibited peaks at around 400 nm (Figure 7-C), and 100 nm

AuNP-streptomycin sulfate mixtures exhibited peaks at around 300 nm (Figure 7-D). It was also

observed that 100 µM streptomycin sulfate in the presence of 60 nm AuNPs and 80 nm AuNPs

presented a peak almost identical to pure AuNPs (Figure 7-B, C). This is likely due to the

formation of aggregates that precipitated out of solution.

It was observed that the 40 nm AuNP-100 µM streptomycin sulfate mixture exhibited

three peaks (Figure 7-A). The exact cause of this is unknown. One possibility is that random

aggregation occurred. A more likely explanation is that an aggregate was formed that was

similar to rod-like aggregates.

The UV-Vis absorption data displayed no peaks for most samples of streptomycin sulfate

(Figure 8). The absence of peaks was caused by the formation of aggregates which then

precipitated out of solution. Similar to the DLS measurements, it was observed that 100 µM

concentrations of streptomycin sulfate mixed with 60 nm and 80 nm AuNPs displayed peaks

almost identical to pure AuNP (Figure 8-B, C). The DLS measurements and the UV-Vis

absorption spectra data suggests that streptomycin sulfate forms a spherical or sphere like

Page 39: Investigation of Aminoglycoside Induced Nanoparticle Self

28

aggregate in water. These aggregates were able to consistently form and be detected at

concentrations above 100 µM.

Page 40: Investigation of Aminoglycoside Induced Nanoparticle Self

29

Figure 7. Intensity-averaged size distribution curves of AuNP-streptomycin sulfate mixtures in nanopure water solvent. Most AuNP-streptomycin sulfate mixtures exhibited one peak at all concentrations above 100 µM. 100 µM streptomycin sulfate didn’t form any detectable aggregates in the presence of 60 nm (B) and 80 nm (C) AuNPs. The 40 nm AuNP-100 µM streptomycin sulfate mixture (A) displayed three peaks. The 100 nm AuNPs (D) detected peaks similar to other concentrations.

Page 41: Investigation of Aminoglycoside Induced Nanoparticle Self

30

Figure 8. UV-Vis absorption spectra of AuNP-streptomycin sulfate mixtures in nanopure water solvent. Most samples did not display any peaks indicating that aggregates formed but then precipitated out of the solution. The 60 nm AuNP (B) and 80 nm AuNPs (C) with 100 µM streptomycin sulfate formed peaks similar to the control.

Page 42: Investigation of Aminoglycoside Induced Nanoparticle Self

31

Neomycin Sulfate in Nanopure Water

Analysis of neomycin sulfate, mixed with different sized AuNPs at different

concentrations of neomycin, using the DLS assay consistently presented one strong peak (Figure

9). It was observed that this single peak varied in size depending on the size of the AuNP

solution mixed with the neomycin sulfate: (A) 40 nm AuNP-neomycin sulfate solutions had Z

average diameters close to 700 nm (Figure 9-A), (B) 60 nm AuNP-neomycin sulfate mixtures had

Z average diameters of approximately 600 nm (Figure 9-B), (C) 80 nm AuNP-neomycin sulfate

AuNP mixtures had Z average diameters ranging between 400 nm and 500 nm (Figure 9-C), (D)

100 nm AuNP-neomycin sulfate mixtures had Z average diameters approximately 300 nm

(Figure 9-D). These findings were supported by the UV-Vis absorption data (Figure 10). No

meaningful peaks were detected at any neomycin sulfate concentration with any AuNP

solutions. The absence of peaks was caused by aggregates forming and then precipitating out

of the solution. The DLS assay data and UV-Vis spectra data indicate that neomycin sulfate

forms spherical aggregates in nanopure water at all tested concentrations.

Page 43: Investigation of Aminoglycoside Induced Nanoparticle Self

32

Figure 9. Intensity-averaged size distribution curves of AuNP-neomycin sulfate mixtures in nanopure water solvent. Neomycin sulfate exhibited one peak at all concentrations with all AuNP sizes used.

Page 44: Investigation of Aminoglycoside Induced Nanoparticle Self

33

Figure 10. UV-Vis absorption spectra of AuNP-neomycin sulfate mixtures in nanopure water. Results are consistent with DLS data. No peaks were observed due to aggregates forming and then precipitating out of the solution.

Page 45: Investigation of Aminoglycoside Induced Nanoparticle Self

34

Neomycin Sulfate at Lower Concentrations in Nanopure Water

It was observed that neomycin sulfate normally displayed one peak, indicating that it

formed a spherical aggregate at concentrations between 1 mM and 100 µM (Figure 11).

Testing was done at lower neomycin sulfate concentrations to see if neomycin sulfate was able

to form aggregates at lower concentrations that could be detected by the DLS assay and UV-Vis

absorption spectroscopy. It was observed 40 nm AuNP-neomycin mixtures exhibited one peak

at around 632 nm. This indicated that neomycin sulfate retained the ability to form spherical

aggregates that could be detected by the DLS at concentrations as low as 10 µM (Figure 11-A).

Some anomalies were observed. DLS analysis of the 40 nm AuNP-10 µM neomycin displayed a

weaker second peak that only appeared in two trials.

All mixtures of 60 nm AuNPs and neomycin sulfate exhibited two peaks (Figure 11-B). It

was observed that the 60 nm AuNP-50 µM neomycin sulfate mixture weakly exhibited two

peaks. The 60 nm AuNP-20 µM neomycin sulfate and 60 nm AuNP-10 µM neomycin sulfate

mixtures displayed two prominent peaks. The 80 nm and 100 nm AuNP mixtures did not

exhibit any peak shifts except for the 50 µM neomycin sulfate mixtures. DLS analysis of the 80

nm and 100 nm AuNP solutions mixed with 50 µM neomycin sulfate exhibited one peak with a Z

average diameter that was approximately 390 nm (Figure 11-C and 11-D).

In the UV-Vis absorption spectra data (Figure 12), the AuNP-40 nm and 20 µM neomycin

sulfate mixture solution displayed a prominent second peak. This is in contrast to the DLS

results which did not detect any second peak. This inconsistency may have been caused by the

DLS laser not hitting any AuNP-neomycin sulfate complexes because only a few complexes were

Page 46: Investigation of Aminoglycoside Induced Nanoparticle Self

35

able to form. All 60 nm AuNP mixtures displayed two peaks (Figure 12-B). Of the 60 nm AuNP-

neomycin sulfate mixtures, the 60 nm AuNP-10 µM neomycin sulfate mixture displayed a

stronger first peak. This suggests that neomycin sulfate can self assemble into rod-like

aggregates. The 80 nm AuNP-neomycin sulfate and 100 nm AuNP-neomycin sulfate mixtures

exhibited peaks similar to the pure control outside of the 50 µM neomycin samples (Figure 12-

C, D). The 100 nm AuNP-20 µM neomycin sulfate mixture exhibited a weak second peak (Figure

12-D).

The UV-Vis absorption spectra data supports the findings of the DLS assay. At lower

concentrations, neomycin sulfate can still form spherical aggregates. At low concentrations in

the presence of 60 nm AuNPs, it was observed that neomycin sulfate was able to form rod-like

self assemblies.

Page 47: Investigation of Aminoglycoside Induced Nanoparticle Self

36

Figure 11. Intensity-averaged size distribution curves of AuNP-neomycin sulfate mixtures in nanopure water solvent. The 40 nm AuNP-20 µM neomycin sulfate mixture (A) exhibited a weak second peak. The 60 nm AuNP-20 µM neomycin sulfate and 60 nm AuNP-10 µM neomycin sulfate AuNP mixtures (B) exhibited strong second peaks.

Page 48: Investigation of Aminoglycoside Induced Nanoparticle Self

37

Figure 12. Intensity-averaged size distribution curves of AuNP-neomycin sulfate mixtures in nanopure water solvent. The second peak for the 40 nm AuNP-20 µM mixture (A) is much more prominent than exhibited in DLS. All 60 nm AuNP-neomycin sulfate mixtures (B) displayed a second peak. All 80 nm and 100 nm AuNP mixtures with neomycin sulfate (C, D) displayed peaks similar to the control except for the 50 µM neomycin sulfate mixture.

Page 49: Investigation of Aminoglycoside Induced Nanoparticle Self

38

Amikacin in 100 mM PB Buffer

In an additional experiment, amikacin was dissolved in 100 mM PB buffer to see if it

would self assemble into rod-like aggregates in physiological conditions. DLS data revealed that

two peaks were detected in various sample trials (Figure 13). The 40 nm AuNPs-500 µM

amikacin mixture and 60 nm AuNP-100 µM amikacin mixtures both displayed two peaks (Figure

13-A, B). The majority of sample measurements only exhibited only one peak. This observation

may have been due to fewer rod-like aggregates forming in the mixture solution. As a result.

the particles intercepted by the DLS laser had a higher number of spherical aggregates and

fewer rod-like aggregates. The 40 nm AuNP mixtures exhibited a single peak at approximately

615 nm (Figure 13-A). The single peak exhibited by the 60 nm AuNP mixtures was at

approximately 600 nm (Figure 13-B). The single peak exhibited by the 80 nm AuNP mixtures

ranged from approximately 350 nm to 540 nm (Figure 13-C). The single peak exhibited by the

100 nm AuNP mixtures was at approximately 340 nm (Figure 13-D).

The DLS data is further supported by the UV-Vis absorption spectra data (Figure 14). It

was observed that AuNP-60 nm with 100 µM amikacin mixture displayed two low peaks (Figure

14-B). For the other samples, a weak second peak was observed. These observations support

the proposal that although some rod-like aggregation occurred, more spherical aggregates

were formed. These spherical aggregates, however, precipitated out of solution. It may be

concluded that under physiological conditions amikacin forms concentration dependent rod-

Page 50: Investigation of Aminoglycoside Induced Nanoparticle Self

39

like assemblies. It was also observed that amikacin dissolved in PB buffer compared to

amikacin dissolved in nanopure water exhibited decreased formation of rod-like aggregates.

Page 51: Investigation of Aminoglycoside Induced Nanoparticle Self

40

Figure 13. Intensity-averaged size distribution curves of AuNP-amikacin mixtures in PB buffer solvent. Overall one peak was observed with all amikacin-AuNP mixtures. Occasionally two peaks were observed indicating weak rod-like self assembly behavior. Unlike nanopure water, the 80 nm and 100 nm AuNP mixtures were able to detect some aggregation (C, D).

Page 52: Investigation of Aminoglycoside Induced Nanoparticle Self

41

Figure 14. UV-Vis absorption spectra of AuNP-amikacin mixtures in PB buffer solvent. This data supported the results of the DLS assay. Few peaks were observed due to the aggregates precipitating out of solution. There were some weak second peaks observed in the 60 nm AuNP mixture (B).

Page 53: Investigation of Aminoglycoside Induced Nanoparticle Self

42

Ribostamycin Sulfate in 100 mM PB Buffer

The results of experiments with amikacin resemble some of the results of prior research

done by our laboratory with ribostamycin sulfate. Our past research established that

ribostamycin sulfate self-assembled into rod-like aggregates in nanopure water. In an

additional experiment, ribostamycin sulfate was analyzed in 100 µM Buffer to determine

whether it would exhibit similar activity in physiological conditions. Comparing the results of

ribostamycin sulfate dissolved in PB buffer to amikacin dissolved in PB buffer, it was observed

that ribostamycin sulfate displayed a greater tendency to form rod-like assemblies in

physiological conditions. Analysis of multiple ribostamycin sulfate samples using the DLS assay

presented two peaks (Figure 15). It was observed that the 40 nm AuNP-250 µM ribostamycin

sulfate mixture, 40 nm AuNP-100 µM ribostamycin sulfate mixture, 60 nm AuNP-1 mM

ribostamycin sulfate mixture, 60 nm AuNP-500 µM ribostamycin sulfate mixture, and 60 nm

AuNP-250 µM ribostamycin sulfate mixture displayed two peaks (Figure 15-A, B). The second

peaks were not as strong as those displayed by ribostamycin sulfate-AuNP mixture solutions in

nanopure water.

It was further observed that 80 nm AuNP and 100 nm AuNP mixture solutions mixed

with ribostamycin sulfate displayed peaks that were almost identical to the peaks exhibited by

their respective pure AuNP controls (Figure 15-C, D). The 60 nm AuNP with 100 µM

ribostamycin sulfate sample also displayed peaks that were almost identical to the peaks

exhibited by the pure 60 nm AuNP control (Figure 15-B).

Page 54: Investigation of Aminoglycoside Induced Nanoparticle Self

43

The UV-Vis spectra data supports the DLS data (Figure 16). The 40 nm AuNP with 100

µM ribostamycin sulfate mixture solution exhibited two strong peaks (Figure 16-A). The 60 nm

AuNP with 500 µM and 250 µM ribostamycin sulfate solutions exhibited two peaks (Figure 16-

B). As further support of the DLS assay results, UV-Vis absorption spectra of the AuNP-60 nm-

100 µM ribostamycin sulfate mixture, the 80 nm AuNP-ribostamycin sulfate mixtures, and the

100 nm AuNP-ribostamycin sulfate mixtures were almost identical to the pure AuNP control

(Figure 16-B, C, D). The previous experiment performed by our lab analyzed ribostamycin

sulfate with AuNPs and found that the results for the AuNP-80 nm and AuNP-100 nm mixture

solutions were almost identical to the results for the pure AuNP suspensions (Figure 16-C, D).

These results strongly suggest that ribostamycin sulfate forms concentration dependent

rod-like assemblies in physiological conditions. Compared to amikacin, ribostamycin sulfate is

more able to form rod-like assemblies at different concentrations. Three more tested samples

of ribostamycin sulfate exhibited two peaks. It was observed that the 40 nm AuNP with 100 µM

and 250 µM ribostamycin sulfate consistently displayed two peaks unlike the AuNP-amikacin

mixture solutions.

Page 55: Investigation of Aminoglycoside Induced Nanoparticle Self

44

Figure 15. Intensity-averaged size distribution curves for AuNP-ribostamycin sulfate mixtures in PB Buffer solvent. 40 nm AuNP mixture (A) exhibited two peaks at 250 µM and 100 µM ribostamycin sulfate concentrations. It was observed that the 60 nm AuNP mixture (B) exhibited two peaks at all concentrations except for at the 100 µM ribostamycin sulfate concentration.

Page 56: Investigation of Aminoglycoside Induced Nanoparticle Self

45

Figure 16.UV-Vis absorption spectra of AuNP-ribostamycin sulfate mixtures in PB buffer solvent. There is a strong second peak displayed by the 40 nm AuNP-100 µM ribostamycin sulfate mixture (A). There are weak second peaks displayed by the 60 nm AuNP-ribostamycin sulfate mixture (B). Most 80 nm and 100 nm AuNP mixtures with ribostamycin (C, D) resembled the controls.

Page 57: Investigation of Aminoglycoside Induced Nanoparticle Self

46

CONCLUSION

The three aminoglycosides (amikacin, streptomycin sulfate, and neomycin sulfate) were

successfully analyzed using AuNPs and the shapes of the three aminoglycosides were

characterized. Of the three, amikacin was readily able to form concentration-dependent rod-

like self assembles in nanopure water. Neomycin sulfate was found to form spherical

aggregates at concentrations between 1 mM and 100 µM. At 50 µM, 20 µM, and 10 µM

concentrations, neomycin sulfate was able to form concentration dependent rod-like self

assemblies in nanopure water. These rod-like assemblies were only detected using 60 nm

AuNPs. Streptomycin sulfate was found to form spherical aggregates. It was observed that the

spherical aggregates were not consistently formed at 100 µM concentrations.

It was observed that both amikacin and ribostamycin sulfate retained the ability to self

assemble into rod-like aggregates in PB buffer. Compared to nanopure water, amikacin had a

lower tendancy to form rod-like aggregates and was more inclined to self assemble into

spherical aggregates. Ribostamycin sulfate was able to consistently form rod-like self

assemblies at different concentrations. One observed key difference between the observations

using PB buffer and nanopure water as a solvent was the concentrations in which rod-like

aggregates formed. In nanopure water, rod-like self assemblies tended to form at higher

concentrations such as 1 mM or 500 µM. In PB buffer, rod-like self assemblies formed at

concentrations below 500 µM.

It is important to study the behavior of aminoglycosides and other antibiotics in order to

develop better therapeutic drugs. Increasing numbers of bacteria strains are developing

Page 58: Investigation of Aminoglycoside Induced Nanoparticle Self

47

resistance to aminoglycosides. Understanding aminoglycoside behavior can aid in further

understanding aminoglycoside pathways for uptake and the mechanism of action within the

bacteria. This in turn can help circumvent antibiotic resistance. Comprehensive knowledge of

aminoglycosides behavior can aid in understanding aminoglycoside interactions with the

human body. Gaining further insight into these interactions could aid in synthesizing less toxic

aminoglycosides.

Page 59: Investigation of Aminoglycoside Induced Nanoparticle Self

48

LIST OF REFERENCES

1. Gonzalez, L. S., 3rd; Spencer, J. P., Aminoglycosides: a practical review. American family physician 1998, 58 (8), 1811-20. 2. Huth, M. E.; Ricci, A. J.; Cheng, A. G., Mechanisms of aminoglycoside ototoxicity and targets of hair cell protection. International journal of otolaryngology 2011, 2011, 937861. 3. Mingeot-Leclercq, M. P.; Tulkens, P. M., Aminoglycosides: nephrotoxicity. Antimicrobial agents and chemotherapy 1999, 43 (5), 1003-12. 4. Mahdiyoun, S. M.; Kazemian, H.; Ahanjan, M.; Houri, H.; Goudarzi, M., Frequency of Aminoglycoside-Resistance Genes in Methicillin-Resistant Staphylococcus aureus (MRSA) Isolates from Hospitalized Patients. Jundishapur Journal of Microbiology 2016, 9 (8). 5. Mieszawska, A. J.; Mulder, W. J.; Fayad, Z. A.; Cormode, D. P., Multifunctional gold nanoparticles for diagnosis and therapy of disease. Molecular pharmaceutics 2013, 10 (3), 831-47. 6. Jain, P. K.; Lee, K. S.; El-Sayed, I. H.; El-Sayed, M. A., Calculated absorption and scattering properties of gold nanoparticles of different size, shape, and composition: applications in biological imaging and biomedicine. The journal of physical chemistry. B 2006, 110 (14), 7238-48. 7. Bogdanovic, J.; Colon, J.; Baker, C.; Huo, Q., A label-free nanoparticle aggregation assay for protein complex/aggregate detection and study. Analytical Biochemistry 2010, 405 (1), 96-102. 8. Zheng, T.; Li Sip, Y. Y.; Leong, M. B.; Huo, Q., Linear self-assembly formation between gold nanoparticles and aminoglycoside antibiotics. Colloids and Surfaces B: Biointerfaces 2018, 164, 185-191. 9. Waksman, S. A., Streptomycin: Background, Isolation, Properties, and Utilization. Science 1953, 118 (3062), 259-266. 10. Waksman, S. A.; Lechevalier, H. A.; Harris, D. A., NEOMYCIN—PRODUCTION AND ANTIBIOTIC PROPERTIES. Journal of Clinical Investigation 1949, 28 (5 Pt 1), 934-9. 11. Umezawa, H.; Ueda, M.; Maeda, K.; Yagishita, K.; Kondo, S.; Okami, Y.; Utahara, R.; Osato, Y.; Nitta, K.; Takeuchi, T., Production and isolation of a new antibiotic: kanamycin. The Journal of antibiotics 1957, 10 (5), 181-8. 12. Weinstein, M. J.; Luedemann, G. M.; Oden, E. M.; Wagman, G. H.; Rosselet, J. P.; Marquez, J. A.; Coniglio, C. T.; Charney, W.; Herzog, H. L.; Black, J., GENTAMICIN, A NEW ANTIBIOTIC COMPLEX FROM MICROMONOSPORA. Journal of medicinal chemistry 1963, 6, 463-4. 13. Neu, H. C., Tobramycin: An Overview. The Journal of Infectious Diseases 1976, 134 (Supplement_1), S3-S19. 14. Kawaguchi, H., Discovery, chemistry, and activity of amikacin. J Infect Dis 1976, 134 suppl, S242-8. 15. Vicens, Q.; Westhof, E., Molecular recognition of aminoglycoside antibiotics by ribosomal RNA and resistance enzymes: An analysis of x-ray crystal structures. Biopolymers 2003, 70 (1), 42-57. 16. Young, P. G.; Walanj, R.; Lakshmi, V.; Byrnes, L. J.; Metcalf, P.; Baker, E. N.; Vakulenko, S. B.; Smith, C. A., The Crystal Structures of Substrate and Nucleotide Complexes of Enterococcus faecium Aminoglycoside-2′′-Phosphotransferase-IIa [APH(2′′)-IIa] Provide Insights into Substrate Selectivity in the APH(2′′) Subfamily. Journal of Bacteriology 2009, 191 (13), 4133-43. 17. Chittapragada, M.; Roberts, S.; Ham, Y. W., Aminoglycosides: molecular insights on the recognition of RNA and aminoglycoside mimics. Perspectives in medicinal chemistry 2009, 3, 21-37. 18. Taber, H. W.; Mueller, J. P.; Miller, P. F.; Arrow, A. S., Bacterial uptake of aminoglycoside antibiotics. Microbiological Reviews 1987, 51 (4), 439-57.

Page 60: Investigation of Aminoglycoside Induced Nanoparticle Self

49

19. Hancock, R. E.; Farmer, S. W.; Li, Z. S.; Poole, K., Interaction of aminoglycosides with the outer membranes and purified lipopolysaccharide and OmpF porin of Escherichia coli. Antimicrobial agents and chemotherapy 1991, 35 (7), 1309-14. 20. Bryan, L. E.; Kwan, S., Roles of ribosomal binding, membrane potential, and electron transport in bacterial uptake of streptomycin and gentamicin. Antimicrobial agents and chemotherapy 1983, 23 (6), 835-45. 21. Allison, K. R.; Brynildsen, M. P.; Collins, J. J., Metabolite-enabled eradication of bacterial persisters by aminoglycosides. Nature 2011, 473 (7346), 216-20. 22. Muir, M. E.; van Heeswyck, R. S.; Wallace, B. J., Effect of growth rate on streptomycin accumulation by Escherichia coli and Bacillus megaterium. Journal of general microbiology 1984, 130 (8), 2015-22. 23. Bryan, L. E.; Elzen, H., Streptomycin Accumulation in Susceptible and Resistant Strains of Escherichia coli and Pseudomonas aeruginosa. Antimicrobial agents and chemotherapy 1976, 9 (6), 928-38. 24. Hurwitz, C.; Braun, C. B.; Rosano, C. L., Role of ribosome recycling in uptake of dihydrostreptomycin by sensitive and resistant Escherichia coli. Biochimica et Biophysica Acta (BBA) - Nucleic Acids and Protein Synthesis 1981, 652 (1), 168-176. 25. Kotra, L. P.; Haddad, J.; Mobashery, S., Aminoglycosides: Perspectives on Mechanisms of Action and Resistance and Strategies to Counter Resistance. Antimicrobial agents and chemotherapy 2000, 44 (12), 3249-56. 26. François, B.; Russell, R. J. M.; Murray, J. B.; Aboul-ela, F.; Masquida, B.; Vicens, Q.; Westhof, E., Crystal structures of complexes between aminoglycosides and decoding A site oligonucleotides: role of the number of rings and positive charges in the specific binding leading to miscoding. Nucleic Acids Research 2005, 33 (17), 5677-90. 27. Cate, J. H.; Yusupov, M. M.; Yusupova, G. Z.; Earnest, T. N.; Noller, H. F., X-ray crystal structures of 70S ribosome functional complexes. Science 1999, 285 (5436), 2095-104. 28. Michael, K.; Wang, H.; Tor, Y., Enhanced RNA binding of dimerized aminoglycosides. Bioorganic & medicinal chemistry 1999, 7 (7), 1361-71. 29. Montie, T.; Patamasucon, P., Aminoglycosides: the complex problem of antibiotic mechanisms and clinical applications. European journal of clinical microbiology & infectious diseases : official publication of the European Society of Clinical Microbiology 1995, 14 (2), 85-7. 30. Masur, H.; Whelton, P. K.; Whelton, A., Neomycin toxicity revisited. Archives of surgery (Chicago, Ill. : 1960) 1976, 111 (7), 822-5. 31. De Broe, M. E.; Paulus, G. J.; Verpooten, G. A.; Roels, F.; Buyssens, N.; Wedeen, R.; Van Hoof, F.; Tulkens, P. M., Early effects of gentamicin, tobramycin, and amikacin on the human kidney. Kidney international 1984, 25 (4), 643-52. 32. Toubeau, G.; Laurent, G.; Carlier, M. B.; Abid, S.; Maldague, P.; Heuson-Stiennon, J. A.; Tulkens, P. M., Tissue repair in rat kidney cortex after short treatment with aminoglycosides at low doses. A comparative biochemical and morphometric study. Laboratory investigation; a journal of technical methods and pathology 1986, 54 (4), 385-93. 33. Hashino, E.; Shero, M., Endocytosis of aminoglycoside antibiotics in sensory hair cells. Brain research 1995, 704 (1), 135-40. 34. Poole, K., Aminoglycoside Resistance in Pseudomonas aeruginosa. Antimicrobial agents and chemotherapy 2005, 49 (2), 479-87. 35. Ramirez, M. S.; Tolmasky, M. E., Aminoglycoside Modifying Enzymes. Drug resistance updates : reviews and commentaries in antimicrobial and anticancer chemotherapy 2010, 13 (6), 151-171.

Page 61: Investigation of Aminoglycoside Induced Nanoparticle Self

50

36. Edgar, R.; Bibi, E., MdfA, an Escherichia coli multidrug resistance protein with an extraordinarily broad spectrum of drug recognition. Journal of Bacteriology 1997, 179 (7), 2274-80. 37. Karlowsky, J. A.; J Hoban, D.; Zelenitsky, S.; G Zhanel, G., Altered denA and anr gene expression in aminoglycoside adaptive resistance in Pseudomonas aeruginosa. 1997; Vol. 40, p 371-6. 38. Hu, Y.; Liu, A.; Vaudrey, J.; Vaiciunaite, B.; Moigboi, C.; McTavish, S. M.; Kearns, A.; Coates, A., Combinations of beta-lactam or aminoglycoside antibiotics with plectasin are synergistic against methicillin-sensitive and methicillin-resistant Staphylococcus aureus. PloS one 2015, 10 (2), e0117664. 39. Faraday, M., X. The Bakerian Lecture. —Experimental relations of gold (and other metals) to light. Philosophical Transactions of the Royal Society of London 1857, 147, 145-181. 40. Kimling, J.; Maier, M.; Okenve, B.; Kotaidis, V.; Ballot, H.; Plech, A., Turkevich Method for Gold Nanoparticle Synthesis Revisited. The Journal of Physical Chemistry B 2006, 110 (32), 15700-15707. 41. Alkilany, A. M.; Murphy, C. J., Toxicity and cellular uptake of gold nanoparticles: what we have learned so far? Journal of Nanoparticle Research 2010, 12 (7), 2313-2333. 42. Huang, X.; Neretina, S.; El-Sayed, M. A., Gold Nanorods: From Synthesis and Properties to Biological and Biomedical Applications. Advanced Materials 2009, 21 (48), 4880-4910. 43. Chithrani, B. D.; Ghazani, A. A.; Chan, W. C. W., Determining the Size and Shape Dependence of Gold Nanoparticle Uptake into Mammalian Cells. Nano Letters 2006, 6 (4), 662-668. 44. Medley, C. D.; Smith, J. E.; Tang, Z.; Wu, Y.; Bamrungsap, S.; Tan, W., Gold nanoparticle-based colorimetric assay for the direct detection of cancerous cells. Analytical chemistry 2008, 80 (4), 1067-72. 45. Zhang, P.; Lee, S.; Yu, H.; Fang, N.; Ho Kang, S., Super-resolution of fluorescence-free plasmonic nanoparticles using enhanced dark-field illumination based on wavelength-modulation. Scientific Reports 2015, 5, 11447. 46. Zheng, T.; Bott, S.; Huo, Q., Techniques for Accurate Sizing of Gold Nanoparticles Using Dynamic Light Scattering with Particular Application to Chemical and Biological Sensing Based on Aggregate Formation. ACS applied materials & interfaces 2016, 8 (33), 21585-94. 47. Khan, M. S.; Vishakante, G. D.; Siddaramaiah, H., Gold nanoparticles: a paradigm shift in biomedical applications. Advances in colloid and interface science 2013, 199-200, 44-58. 48. DeLong, R. K.; Reynolds, C. M.; Malcolm, Y.; Schaeffer, A.; Severs, T.; Wanekaya, A., Functionalized gold nanoparticles for the binding, stabilization, and delivery of therapeutic DNA, RNA, and other biological macromolecules. Nanotechnology, Science and Applications 2010, 3, 53-63. 49. Jazayeri, M. H.; Amani, H.; Pourfatollah, A. A.; Pazoki-Toroudi, H.; Sedighimoghaddam, B., Various methods of gold nanoparticles (GNPs) conjugation to antibodies. Sensing and Bio-Sensing Research 2016, 9 (Supplement C), 17-22. 50. Maeda, H., Tumor-Selective Delivery of Macromolecular Drugs via the EPR Effect: Background and Future Prospects. Bioconjugate Chemistry 2010, 21 (5), 797-802. 51. Dreaden, E. C.; Austin, L. A.; Mackey, M. A.; El-Sayed, M. A., Size matters: gold nanoparticles in targeted cancer drug delivery. Therapeutic delivery 2012, 3 (4), 457-78. 52. Ghosh, P.; Han, G.; De, M.; Kim, C. K.; Rotello, V. M., Gold nanoparticles in delivery applications. Advanced drug delivery reviews 2008, 60 (11), 1307-15. 53. Jurj, A.; Braicu, C.; Pop, L. A.; Tomuleasa, C.; Gherman, C. D.; Berindan-Neagoe, I., The new era of nanotechnology, an alternative to change cancer treatment. Drug design, development and therapy 2017, 11, 2871-2890. 54. Fratoddi, I.; Venditti, I.; Cametti, C.; Russo, M. V., How toxic are gold nanoparticles? The state-of-the-art. Nano Research 2015, 8 (6), 1771-1799. 55. Yah, C. S., The toxicity of Gold Nanoparticles in relation to their physiochemical properties. Biomed Res-India 2013, 24 (3), 400-413.

Page 62: Investigation of Aminoglycoside Induced Nanoparticle Self

51

56. Hackley, V. A.; Clogston, J. D., Measuring the hydrodynamic size of nanoparticles in aqueous media using batch-mode dynamic light scattering. Methods in molecular biology (Clifton, N.J.) 2011, 697, 35-52. 57. Edward, J. T., Molecular volumes and the Stokes-Einstein equation. Journal of Chemical Education 1970, 47 (4), 261. 58. Jiang, J.; Oberdörster, G.; Biswas, P., Characterization of size, surface charge, and agglomeration state of nanoparticle dispersions for toxicological studies. Journal of Nanoparticle Research 2009, 11 (1), 77-89. 59. Rodríguez-Fernández, J.; Pérez−Juste, J.; Liz−Marzán, L. M.; Lang, P. R., Dynamic Light Scattering of Short Au Rods with Low Aspect Ratios. The Journal of Physical Chemistry C 2007, 111 (13), 5020-5025. 60. Liu, H.; Pierre-Pierre, N.; Huo, Q., Dynamic light scattering for gold nanorod size characterization and study of nanorod–protein interactions. Gold Bulletin 2012, 45 (4), 187-195. 61. Jans, H.; Liu, X.; Austin, L.; Maes, G.; Huo, Q., Dynamic Light Scattering as a Powerful Tool for Gold Nanoparticle Bioconjugation and Biomolecular Binding Studies. Analytical chemistry 2009, 81 (22), 9425-9432.