20.01.2013 Views

ABSTRACT ROSKOV, KRISTEN EKIERT. Engineered ...

ABSTRACT ROSKOV, KRISTEN EKIERT. Engineered ...

ABSTRACT ROSKOV, KRISTEN EKIERT. Engineered ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>ABSTRACT</strong><br />

<strong>ROSKOV</strong>, <strong>KRISTEN</strong> <strong>EKIERT</strong>. <strong>Engineered</strong> Organometallic Polymer and Hybrid Systems<br />

Containing Nanoparticles and/or Poly(ferrocenylsilanes). (Under the direction of Professor<br />

Richard Spontak.)<br />

Formation of polymer nanocomposites is becoming an increasingly attractive and<br />

facile means by which to combine the desirable properties of metals and metal oxides (e.g.,<br />

electrical, magnetic, optical, and thermal) with those of polymers (e.g., flexible, lightweight<br />

and tough). Incorporation of nanoscale objects such as spheroidal nanoparticles or elongated<br />

nanorods into electrospun polymer nano/microfibers measuring from 50 nm to 1 μm in<br />

diameter yields functional nanomaterials that can be used in various applications ranging<br />

from data storage and conductive nanowires to nonwoven sensors, magnetic filters and drug<br />

delivery patches. By aligning nanoscale objects in one-dimensional constructs, we expect<br />

that desirable attributes arising from highly anisotropic electronic, optical, thermal, magnetic,<br />

and catalytic properties can be realized. The objective of this study is to gain a better<br />

fundamental understanding of how to controllably align and position nanoparticles and<br />

nanorods within polymer nano/microfibers to generate unique properties. To achieve this<br />

objective, we focus on four specific process strategies. In the first, superparamagnetic iron<br />

oxide nanoparticles (SPIONs) are aligned into one-dimensional nanoarrays through the use<br />

of magnetic field-assisted electrospinning. In this case, an electromagnet is positioned near<br />

the Taylor cone of the suspension to be electrospun so that the magnetic field is oriented<br />

perpendicular to the electric field. Transmission electron microscopy (TEM) is utilized to<br />

ascertain the morphology of the resultant nanocomposite fibers and reveals that the SPION


nanoarrays persist intact beyond 1 μm. Since the magnetic field can be pulsed, the length of<br />

the nanoarrays can be judiciously controlled. Magnetization hysteresis curves measured on a<br />

superconducting quantum interference device yield saturation magnetization and mean<br />

magnetic moment values. Secondly, gold nanorods (GNRs) varying in aspect ratio have been<br />

flow-aligned in electrospun fibers, and the fibers have likewise been aligned to permit long-<br />

range orientation order at both the nanoscale and macroscale. This is an important<br />

consideration in the fabrication of devices spanning multiple size scales. The GNRs within<br />

nano/microfibers exhibit excellent alignment with their longitudinal axis parallel to the fiber<br />

axis. Optical absorbance spectroscopy measurements reveal that the longitudinal surface<br />

plasmon resonance bands of the aligned GNRs are highly anisotropic, depending on<br />

polarization angle, and that maximum absorption occurs when polarization is parallel to the<br />

fiber axis. Lastly, blends of hydrophobic and hydrophilic polymers have been prepared to<br />

control the spatial position of SPIONs within electrospun fibers on the basis of<br />

thermodynamic compatibility. In this case, TEM confirms that a core-sheath nanostructure<br />

naturally forms due to polymer-polymer phase separation and that the hydrophobic<br />

nanoparticles are sequestered in one preferred phase. Lastly, a nanocomposite fiber is<br />

created using only one entity, the organometallic polymer poly(ferrocenylsilane) (PFS)and its<br />

crystalline structure is probed alone and in the presence of SPIONs. Block copolymer<br />

cylindrical micelles of PFS-b-poly(isoprene) (PI) are crosslinked within an elastomeric<br />

matrix of poly (vinylmethoxysilane) (PVMS) and found to maintain their crystalline structure<br />

with the target application being nanowires in soft electronics.


© Copyright 2011 by Kristen Ekiert Roskov<br />

All Rights Reserved


<strong>Engineered</strong> Organometallic Polymer and Hybrid Systems Containing Nanoparticles and/or<br />

Poly(ferrocenylsilanes)<br />

by<br />

Kristen Ekiert Roskov<br />

A dissertation or submitted to the Graduate Faculty of<br />

North Carolina State University<br />

in partial fulfillment of the<br />

requirements for the degree of<br />

Doctor of Philosophy<br />

Chemical and Biomolecular Engineering<br />

Raleigh, North Carolina<br />

2011<br />

APPROVED BY:<br />

_______________________________ ______________________________<br />

Richard Spontak Saad Khan<br />

Committee Chair<br />

________________________________ ________________________________<br />

Michael Dickey Russell Gorga


DEDICATION<br />

To my parents, Andrea and Bernie, for their steadfast love and support. You have taught me<br />

life’s greatest lessons by example and that has shaped the woman I have become. I love you<br />

so much.<br />

ii


BIOGRAPHY<br />

Kristen was born in Pittsburgh, PA in 1983 to Andrea and Bernie and is an only child.<br />

After graduating from North Allegheny High School, she went on to the University of<br />

Maryland and received a B.S. in Chemical Engineering in 2006. In the summer of 2005, she<br />

was awarded a summer undergraduate research fellowship at the National Institute of<br />

Standards and Technology in Gaithersburg, MD and was hired as an undergraduate research<br />

assistant for the following year. This research experience at NIST resulted in four<br />

publications and was the experience that sparked her interest in graduate school. From there,<br />

she moved to North Carolina to attend graduate school in the department of Chemical and<br />

Biomolecular Engineering at North Carolina State University. In 2007, Kristen was awarded<br />

the National Science Foundation graduate research fellowship and in 2010 she was awarded<br />

the North Carolina Space Grant Fellowship. Kristen is enthusiastic about mentoring and<br />

tutoring girls and encouraging them to pursue a career in science. Kristen will be joining the<br />

Ph.D. Professional Development Program with BASF upon graduation.<br />

iii


ACKNOWLEDGMENTS<br />

Foremost, I’d like to thank my advisor, Dr. Richard Spontak. You have taught me to<br />

be the scientist I am today. The biggest lesson I have learned from you is that the outcome is<br />

usually not what you expect, but discovering the ‘why’ is the best part. You see your<br />

students in the best possible light and I have always strived to meet those expectations.<br />

Thank you for everything.<br />

I’d like to thank my mentors at NIST who motivated me to attend graduate school<br />

and initially sparked my interest in research, namely Thomas Epps, Christopher Stafford,<br />

Michael Fasolka, and Matthew Becker. I can truly say that you shaped the direction of my<br />

life and career. Thank you for taking the time to invest in a SURF student!<br />

My thesis committee, Saad Khan, Michael Dickey, and Russell Gorga, have been<br />

extremely helpful and supportive. I’d also like to thank many other faculty at NCSU; Jan<br />

Genzer, Saad Khan, Joe Tracy, Kirill Efimenko, Roberto Garcia, and Chuck Mooney. I have<br />

learned so much from all of you and thank you for all your help and fruitful discussions we<br />

have had. Thank you to the professors at other universities with whom I collaborated with;<br />

specifically Ian Manners, Lyudmila Bronstein, and Amy Oldenburg.<br />

My fellow PMG peers and officemates made the day-to-day life of graduate school<br />

much more interesting and I feel very lucky to have made lifelong friends. I’d like to<br />

specifically thank Arjun Krishnan, Pruthesh Varagantwar, Omer Gozen, Evren Ozcam, and<br />

Anand Patel. To Arjun and Anand- I can’t imagine going through this experience without<br />

either of you. Thank you for always answering my deluge of questions and lending a helping<br />

iv


hand. Thank you to the undergraduate students that helped with this research; Raleigh Davis<br />

and Kathryn Earley. To Sara for being my constant throughout this journey-I am so grateful<br />

that we met and decided to live together at Cardinal Club! You have motivated me to be a<br />

true friend, a better researcher, and a great cook. To Michelle, for constantly being there on<br />

so many different levels and always listening. To Julie, for enthusiastically reading every<br />

publication and always asking how my polymers are doing.<br />

I’d like to thank all the support I have received from not only my parents but many<br />

other relatives. My grandma Helen Roskov was a strong woman with so much love for her<br />

family. I thank her for the unwavering encouragement she always gave me and wish she<br />

could be here to see what I’ve achieved. I am so thankful for my ‘sister’ Lisa, who initially<br />

suggested I become an engineer when I was young to which I emphatically replied ‘I don’t<br />

want to drive trains!’ You are such a giving, thoughtful person and I’ve always strived to be<br />

half as caring as you. To my goddaughter Sienna, I love you so much and am so excited to<br />

see you grow. I have been so fortunate to have my godparents Aunt Dee Dee and my Uncle<br />

Kenny. It’s not the same without you being here, we miss you terribly. Thank you to all my<br />

other friends and family for being there for me during this journey. I am incredibly lucky.<br />

v


TABLE OF CONTENTS<br />

LIST OF TABLES ................................................................................................................... ix<br />

LIST OF FIGURES .................................................................................................................. x<br />

CHAPTER I: Introduction and Overview ............................................................................... 1<br />

1.1 Introduction ........................................................................................................................ 1<br />

1.1.1 Electrospinning ................................................................................................... 2<br />

1.1.2 Control over Fiber Characteristics ...................................................................... 6<br />

1.1.3 Electropsinning Fiber Material ......................................................................... 12<br />

1.1.4 Imparting Functionality to Electropsun Fibers ................................................. 18<br />

1.1.5 Applications ...................................................................................................... 23<br />

1.2 One-Dimensional Magnetic Nanostructures .................................................................... 28<br />

1.3 Organometallic Polymers ................................................................................................. 31<br />

1.3.1 Overview ........................................................................................................... 31<br />

1.3.2 Poly(ferrocenylsilanes) ...................................................................................... 33<br />

Nomenclature .......................................................................................................................... 38<br />

Figures..................................................................................................................................... 39<br />

References ............................................................................................................................... 61<br />

CHAPTER II: Long-Range Alignment of Gold Nanorods in Electrospun Polymer Nano/<br />

Microfiber ...................................................................................................................... 76<br />

Figures..................................................................................................................................... 87<br />

References ............................................................................................................................... 91<br />

CHAPTER III: Magnetic Field-Induced Alignment of Nanoparticles in Electrospun<br />

Microfibers ...................................................................................................................... 94<br />

Figures................................................................................................................................... 106<br />

References ............................................................................................................................. 113<br />

CHAPTER IV: Using Polymer Blend Morphology to Position Ligand-Functionalized<br />

Nanoparticles in Electrospun Polymer Microfibers .............................................................. 116<br />

Tables ................................................................................................................................... 133<br />

Figures................................................................................................................................... 135<br />

References ............................................................................................................................. 142<br />

CHAPTER V: Nanostructured Organometallic Polymer Systems Containing<br />

Poly(ferrocenylsilanes) ......................................................................................................... 145<br />

5.1 Introduction .................................................................................................................... 145<br />

5.2 Experimental .................................................................................................................. 149<br />

5.2.1 Materials .......................................................................................................... 149<br />

5.2.2 Synthesis of Specialty Polymers ..................................................................... 149<br />

5.2.3 Preparation of Nanoparticles........................................................................... 150<br />

vi


5.2.4 Preparation of Electrospun Fibers ................................................................... 151<br />

5.2.5 Characterization of PFS Nanomaterials .......................................................... 151<br />

5.3 Electrospinning and Characterization of PFS Homopolymers ...................................... 151<br />

5.4 Phase behavior of binary blends of PFS in Elastomeric Matrices ................................. 156<br />

5.4.1 PFDMS-b-PI/PI Blend by Solvent Casting ..................................................... 158<br />

5.4.2 Cross-linking within Polyisoprene .................................................................. 158<br />

5.4.3 Shell-Cross-Linking of Cylindrical PI-b-PFS Micelles .................................. 161<br />

5.4.4 Cross-linked PVMS as the Matrix Polymer.................................................... 162<br />

5.5 Conclusion ..................................................................................................................... 165<br />

Tables ................................................................................................................................... 166<br />

Figures................................................................................................................................... 171<br />

References ............................................................................................................................. 185<br />

CHAPTER VI: Conclusions and Future Work ..................................................................... 189<br />

6.1 Conclusions .................................................................................................................... 189<br />

6.1.1 Long-Range Alignment of Gold Nanorods in Electrospun Polymer<br />

Nano/Microfibers ........................................................................................... 190<br />

6.1.2 Magnetic Field-Induced Alignment of Nanoparticles in Electrospun<br />

Microfibers ..................................................................................................... 190<br />

6.1.3 Using Polymer Blend Morphology to Position Ligand-Functionalized<br />

Nanoparticles in Electrospun Polymer Microfibers ....................................... 191<br />

6.1.4 Nanostructured Organometallic Polymer Systems Containing<br />

Poly(ferrocenylsilanes) ................................................................................... 191<br />

6.2 Recommendations for future work ................................................................................. 192<br />

6.2.1 Gold Nanorod Alignment through Electrospun Fiber Degradation ............... 192<br />

6.2.2 Field Uniformity in Magnetic-Assisted Electrospinning ............................... 193<br />

6.2.3 Poly(ferrocenylsilane) Cylindrical Micelles Oriented within Electropun<br />

Fibers .............................................................................................................. 194<br />

Figures................................................................................................................................... 197<br />

References ............................................................................................................................. 201<br />

APPENDIX ........................................................................................................................... 203<br />

APPENDIX I: Responsive PET Nano/Microfibers via Surface-Initiated<br />

Polymerization .......................................................................................................... 203<br />

APPENDIX II: Generation of Functional PET Microfibers through Surface-Initiated<br />

Polymerization .......................................................................................................... 224<br />

APPENDIX III: Modification of Melt-spun Isotactic Polypropylene and Poly(lactic acid)<br />

Bicomponent Filaments with a Premade Block Copolymer ..................................... 261<br />

APPENDIX IV: Enhanced Biomimetic Performance of Ionic Polymer-Metal Composite<br />

Actuators Prepared with Nanostructured Block Ionomers ....................................... 307<br />

vii


APPENDIX V: Block Copolymer Self-Organization vs. Interfacial Modificationin Bilayered<br />

Thin-Film Laminates ................................................................................................ 329<br />

viii


LIST OF TABLES<br />

Table 4.1. XRD characteristics of PEO powder and electrospun PEO/P2VP<br />

microfibers. ................................................................................................... 133<br />

Table 4.2. XRD characteristics of electrospun PEO/P2VP microfibers with SPIONs<br />

measuring 18 nm in diameter and added at a concentration of 2.5 vol%. .... 134<br />

Table 5.1. Characteristics of the (Co)Polymers Used in this Study. .............................. 166<br />

Table 5.2. PFDMS Fiber Diameters............................................................................... 167<br />

Table 5.3. X-ray diffraction bragg angle, d-spacing, and average crystallite size for 96<br />

kDa poly(ferryocenydimethylsilane) powder and electrospun fibers. .......... 168<br />

Table 5.4. Components utilized in the vulcanization of poly(isoprene). ....................... 169<br />

Table 5.5. Solubility parameters of polymers relative to the solvent n-hexane. ............ 170<br />

ix


LIST OF FIGURES<br />

Figure 1.1. Schematic figure of basic electrospinning setup .............................................. 39<br />

Figure 1.2. Photograph of the whipping motion of the instability region of the polymer jet<br />

during the electrospinning process. .................................................................. 40<br />

Figure 1.3. A method to produce aligned fibers. A) a schematic utilizing the parallel<br />

grounded electrode collection system and B) the resulting aligned polymer<br />

fibers ................................................................................................................. 41<br />

Figure 1.4. SEM micrograph of beaded PEO fibers. .......................................................... 42<br />

Figure 1.5. Reduction in bead density upon increase in polymer solution viscosity .......... 43<br />

Figure 1.6. Ribbon-like fibers formed via electrospinning ................................................. 44<br />

Figure 1.7. SEM image of anatase hollow fibers created via a co-axial electrospinning<br />

setup.................................................................................................................. 45<br />

Figure 1.8. TEM of PPX/Pd TUFT hybrid nanotubes after the pyrolysis of PLA template<br />

fibers and inset is an electron diffraction pattern of Pd crystals ...................... 46<br />

Figure 1.9. SEM micrograph of porous PLA fibers obtained via electrospinning and<br />

subsequent swelling .......................................................................................... 47<br />

Figure 1.10. SEM iamges of anatase nanofibers whose surfaces have been decorated with a)<br />

gold and b) silver nanoparticles via photocatalytic reduction .......................... 48<br />

Figure 1.11. A list of organosoluble polymers and their molecular structure ...................... 49<br />

Figure 1.12. Schematic of polymer/inorganic composite nanofibers when a) inorganic ions<br />

are incorporated into electrospun fibers followed by exposure to gas to<br />

synthesize inorganic nanoparticles both inside and outside of the nanofiber and<br />

b) when only the surface of nanofibers are modified with metal ions. ............ 52<br />

Figure 1.13. TEM micrographs of multicomponent polymer electrospun fibers<br />

demonstrating a) A core-sheath structure formed by a polymer blend b)<br />

Lamellar structure formed by a phase-separated block copolymer c) Cylindrical<br />

structure formed by a phase-separated block copolymer ................................. 53<br />

Figure 1.14. TEM image of a) PAN/CNT composite nanofiber mat and b) demonstrating the<br />

uniform distribution and alignment of CNTs witin a PAN fiber ..................... 54<br />

Figure 1.15. Average Young’s modulus for electrosopun nylon-6 and nylon-6/O-MMT<br />

nanocomposite single fibers vs. fiber diameter ................................................ 55<br />

Figure 1.16. Demonstration that as the aspect ratio of gold nanorods increases, as does the<br />

maximum optical absorbance and thus the color of the aqueous colloidal<br />

suspension. ....................................................................................................... 56<br />

Figure 1.17. Three-dimensional mineralized electrospun fibers mimicking the hierarchical<br />

structure of bone ............................................................................................... 57<br />

Figure 1.18. Left: magnetic induction map from two pairs of bacterial magnetite chains.<br />

Right: A bright-field TEM image of a double chain of magnetite<br />

magnetosomes. ................................................................................................. 58<br />

x


Figure 1.19. Pyroloysis of UV cross-linked PS-b-PFEMS films with a) height-mode<br />

scanning force microscopy, b) phase-mode, c) TEM images, and d) A<br />

schematic of the morphology. Inset scale bars = 50 nm. ................................ 59<br />

Figure 1.20. TEM micrographs of scarf-shaped PI-b-PFS co-micelles (scale bar<br />

= 500 nm) ......................................................................................................... 60<br />

Figure 2.1. TEM image of GNRs deposited from an aqueous suspension onto a carboncoated<br />

TEM grid. The inset shows the distribution of measured aspect ratios of<br />

the GNRs, which measure 49 nm long and 17 nm in diameter on<br />

average.............................................................................................................. 87<br />

Figure 2.2. SEM image of macroscopically-aligned electrospun PEO fibers containing<br />

GNRs. ............................................................................................................... 88<br />

Figure 2.3. Aligned GNRs in electrospun PEO nano/microfibers as functions of fiber<br />

diameter and GNR volume fraction (��: (a) 40 nm and (�� = 0.006, (b) 50 nm<br />

and (�� = 0.045, (c) 650 nm and (���= 0.035, and (d) 3000 nm and (��= 0.031.<br />

A selected-area electron diffraction pattern of the corresponding sample is<br />

included as an inset in (b). ................................................................................ 89<br />

Figure 2.4. Absorbance spectra for (a) randomly oriented GNRs in a PEO film measuring<br />

~500 �m thick at different polarization angles and (b) GNRs aligned within<br />

electrospun PEOmicrofibers measuring ~200 nm in diameter at polarization<br />

angles varying from 0° (parallel to the fiber axis n) to 90° (perpendicular to n).<br />

In both cases, the data are color-coded and labeled in (a). ............................... 90<br />

Figure 3.1. Schematic illustration of the magnetic field-assisted electrospinning setup used<br />

in this study. Note the position of the electromagnet, which yields a magnetic<br />

field that is perpendicular to the electric field employed during<br />

electrospinning. .............................................................................................. 106<br />

Figure 3.2. TEM images of randomly dispersed SPIONs in electrospun PCL microfibers<br />

varying in SPION concentration (in vol%): (a) 0.5 and (b) 2.5. A TEM image<br />

of SPIONs measuring 17.6 nm in diameter and drop cast from chloroform is<br />

included in the inset of (a). The scalemarker in the inset corresponds to<br />

50 nm. ............................................................................................................. 107<br />

Figure 3.3. TEM images of magnetic field-aligned SPIONs, measuring 17.6 nm in<br />

diameter, in PCL microfibers illustrating long, contiguous arrays in (a), and<br />

shorter, pulsed arrays in (b). The scalemarker in the inset corresponds to 100<br />

nm ................................................................................................................. 108<br />

xi


Figure 3.4. Magnetization (M) hysteresis curves at 300 K as a function of the magnetizing<br />

field strength (H) for SPIONs measuring 17.6 nm in diameter. In (a), the<br />

hysteresis curves are measured from unembedded SPIONs ( ), as well as<br />

randomly dispersed and magnetically aligned SPIONs in electrospun PCL<br />

microfibers (blue and red, respectively). The inset shows magnetization<br />

hysteresis curves recorded for the embedded SPIONS at low fields and ambient<br />

temperature. In (b), the hysteresis curves from the SPIONs embedded in PCL<br />

microfibers (see the corresponding diagrams) are fitted to Eq. 2 in the text<br />

(solid lines) to discern the saturation magnetization and mean dipole moment<br />

from each dataset. ........................................................................................... 109<br />

Figure 3.S1. SEM image of SPION-containing PCL fibers electrospun in the presence of an<br />

external magnetic field. The inset shows evidence of surface dimpling on a<br />

large fiber. The scalemarker in the inset corresponds to 2 �m. ..................... 111<br />

Figure 3.S2. EDS spectrum of a SPION-containing PCL fiber electrospun in the presence of<br />

an external magnetic field. The elements responsible for the observed peaks are<br />

labeled, and the x-ray energies associated with the K� and L lines of Fe are<br />

identified by the blue lines. ............................................................................ 112<br />

Figure 4.1. TEM image of SPIONs measuring 16.4 nm in diameter and drop cast from<br />

chloroform. ..................................................................................................... 135<br />

Figure 4.2. (a) SEM image of SPION-containing PEO/P2VP microfibers composed of 80<br />

wt% PEO and electrospun from an 8.5 wt% solution in chloroform. (b) An<br />

enlargement showing the surface of the microfibers included in (a). The inset<br />

in (b) displays a SPION-rich bead, the scalemarker corresponds to<br />

500 nm. ........................................................................................................... 136<br />

Figure 4.3. TEM images of SPIONs measuring 16.4 nm in diameter and dropcast from<br />

chlorform....................................................................................................... 138<br />

Figure 4.4. XRD patterns acquired from PEO powder and electrospun microfibers<br />

composed of PEO/P2VP at different PEO concentrations (labeled). ........... 138<br />

Figure 4.5. XRD patterns acquired from electrospun microfibers composed of PEO/P2VP<br />

with SPIONs (18 nm and 2.5 vol%) at different PEO concentrations<br />

(labeled). ....................................................................................................... 139<br />

Figure 4.6. Average PEO crystal size (t) extracted from XRD patterns and presented as a<br />

function of PEO concentration parallel (circles) and perpendicular (triangles)<br />

to the fiber axis for systems without (open symbols) and with (filled symbols)<br />

SPIONs. Values measured from PEO powder are included (triangles). The<br />

solid and dashed lines serve as guides for the eye. ....................................... 140<br />

xii


Figure 4.7. Schematic illustration depicting the arrangement of polymer chains in a<br />

core/sheath microstructure of PEO/P2VP (a) before and (b) after SPION<br />

addition (not to scale). Addition of SPIONs promotes a reduction in crystal<br />

size but a more parallel chain arrangement with respect to the fiber axis. ... 141<br />

Figure 5.1. Schematic of PFEMS synthesis in which the noted molecules are a)<br />

ferrocenophane b) ethylmethylsilaferrocenophane and c) poly<br />

(ferrocenylethylmethylsilane). ...................................................................... 171<br />

Figure 5.2. SEM micrographs of a) 15% PFDMS in THF:DMF without surfactant b) 15%<br />

PFDMS in THF:DMF with surfactant c) 20% PFDMS in DCM with<br />

surfactant and d) 18% PFPMS in THF:DMF with surfactant. All scale bars<br />

represent 20 μm. ............................................................................................ 172<br />

Figure 5.3. PS-b-PFS-b-P2VP lithographic template used for the preparationof Nanoscale<br />

magnetic dots. a) Phase-separation of the triblock in the bulk. b) Hollow PFS<br />

cylinders are formed after etching because it has a selective resistance. c)<br />

Profile of the hollow PFS cylinders. (Used with permission by Jessica<br />

Gwyther from Bristol University.) ................................................................ 173<br />

Figure 5.4. XRD curves from 2Ɵ = 7 – 25 °for PFDMS powder, fibers, and fibers with<br />

both 10 and 14 nm iron oxide nanoparticles. ................................................ 174<br />

Figure 5.5. Schematic demonstrating the chains of PFDMS and corresponding d-spacing<br />

between adjacent ferrocene units in a) powder form, b) in electrospun fibers,<br />

c) in electrospun fibers with larger (~14 nm) iron oxide nanoparticles, and d)<br />

in electrospun fibers with smaller (~10 nm) iron oxide nanoparticles. ........ 174<br />

Figure 5.6. TEM micrographs of PFS-b-PI micelles dropcast from a 1 mg/mL hexane<br />

solution onto a carbon-coated TEM grid with an average width of 14.9 nm and<br />

lengths exceeding one micron. ...................................................................... 176<br />

Figure 5.7. TEM micrographs of PFS-b-PI micelles blended at a ratio of a) 3:100 with PI<br />

and b) 3:25 with PI. ....................................................................................... 178<br />

Figure 5.8. Schematic of the vulcanization and micelle crosslinking technique. ........... 178<br />

Figure 5.9. TEM micrographs of 1:1000 PFS-b-PI:PI vulcanized at ~120 °C for 5 housr<br />

demonstrating a complete dissolution of the micelles with only iron<br />

nanoparticles remaining. ............................................................................... 179<br />

Figure 5.10. TEM micrographs of shell cross-linked PFS-b-PI with an average width of 43<br />

nm and lengths exceeding 1.5 microns dropcast from a hexane solution. .... 180<br />

Figure 5.11. TEM micrographs of PFS-b-PI micelles blended with a 5 wt% PI solution in<br />

hexane and dropcast onto a carbon-coated TEM grid. ................................. 181<br />

Figure 5.12. TEM micrographs of microtomed PFS-b-PI micelles blended with PVMS at<br />

ratios of 1:300 and 1:600 at thicknesses of ~120 nm. ................................... 182<br />

xiii


Figure 5.13. TEM micrographs shell cross-linked PFS-b-PI dropcast from a hexane<br />

solution as a) a control, b) heated to 70 °C for 30 minutes and c) stirred for<br />

several minutes mimicking conditions during the cross-linking of the PVMS<br />

solution. ......................................................................................................... 182<br />

Figure 5.14. SEM micrographs of a PVMS cross-linked film containing PFS-b-PI micelles<br />

a) looking down the surface from the cross-section and b) of the fractured<br />

cross-section where the scalebar of the inset refers to 200 μm. .................... 182<br />

Figure 6.1. Dropcast PFDMS-b-P2VP cylindrical micelles onto a carbon coated TEM<br />

grid from DMF. ............................................................................................. 182<br />

Figure 6.2. SEM micrographs of 32 wt% P2VP fibers electrospun from 9:1 DMF:THF a)<br />

without micelle addition and b) with PFDMS-b-P2VP micelles. ................. 182<br />

Figure 6.3. TEM micrograph of 32 wt% P2VP fibers electrospun with PFDMS-b-P2VP<br />

micelles. ........................................................................................................ 182<br />

xiv


Introduction<br />

Chapter I<br />

Introduction and Overview<br />

As technology has developed in the last century, it has become possible to investigate<br />

matter on a smaller and smaller scale and to manipulate matter on an atomic and molecular<br />

scale. This development has led to the creation of the buzz word “nanotechnology.” which<br />

typically refers to structures with a size scale between 1 and 100 nanometers. If we<br />

specifically look at the construction of soft materials, they can be created through two<br />

different approaches: a “bottom-up” approach, where molecules self-assemble on the atomic<br />

level; and a “top-down” approach, using patterned templates to achieve atomic order. Soft<br />

nanomaterials viewed through either approach are interesting mainly due to the weak, non-<br />

covalent bonding that occurs within them and the ability to use thermal energy to break and<br />

re-form these bonds. Changes in morphology can be induced by thermal energy such as<br />

temperature, pH, or other external triggers. The term “nanocomposite” refers to a multiphase<br />

material where one of the phases is nano-scaled and the phases differ both in structure and<br />

chemistry, thereby allowing the formation of multifunctional materials with very high surface<br />

area-to volume ratios of the dispersed phase. In this chapter, the creation of polymeric<br />

1


materials with functional properties by utilizing electrospinning as a fabrication tool and the<br />

processing or organometallics will be discussed.<br />

1.1.1 Electrospinning<br />

History<br />

Generally, electrospinning is the process of forming fibers through the application of<br />

electrostatic forces. It is an extension of the electrostatic spray technique where a high<br />

voltage is applied to a liquid jet in order to form small droplets. Electrospinning was initially<br />

investigated by Lord Rayleigh in 1882 1 when he questioned the electric potential that would<br />

be needed to overcome the surface tension of a drop. Then, in 1914, Zeleny et al.<br />

investigated the behavior of fluid droplets at the end of glass capillaries. 2 Electrospray is<br />

used to manufacture particles of varying size and is used in applications including mass<br />

spectrometry, painting, and inkjet printing. 3 The electrospinning of plastics first appeared as<br />

a patent in 1934 by Formhals 4 however it was decades later in the 1990’s when the Reneker<br />

group revived interest in this technique 5 with a goal of forming ultrathin, continuous polymer<br />

fibers. Other fiber forming techniques include drawing, 6 melt-blowing, 7 and multicomponent<br />

fibers. 7 Though these last two in particular have very high productivity, only electrospinning<br />

has the ability to form continuous fibers with nanoscale dimensions and with flexibility in<br />

terms of polymer choice. Since then, the interest in electrospinning has exploded and more<br />

than 100 polymers have been electrospun 8 due to the simplicity of the process and versatility<br />

it affords.<br />

2


An example of an electrospinning scheme is seen in Figure 1.1. First in the<br />

eletrospinning process, a polymer solution is drawn into a syringe with a spinneret, or<br />

metallic needle, and connected to a syringe pump that produces a steady and controllable rate<br />

of discharge. A direct current power supply creates an electric field with high voltages<br />

between the needle tip and grounded collection plate that typically ranges from 1-20 kV.<br />

Surface tension is responsible for holding a droplet of polymer solution on the tip of the<br />

spinneret and when an electrode is connected to the spinneret and a voltage is applied, a<br />

charge develops on the surface of the liquid due to mutual charge repulsion. 9 As the voltage<br />

is increased, the hemispherical shape of the droplet becomes elongated into a conical<br />

structure defined as a “Taylor cone” 10 and as the intensity of the electric field increases<br />

further, a critical threshold is passed where electrostatic forces overcome surface tension and<br />

a charged jet is emitted from the tip of the Taylor cone. The ejected polymer solution<br />

undergoes a whipping process, 11 as seen in Figure 1.2 12 between the spinneret tip and the<br />

collection plate during which time the solvent is evaporated. The dry, highly porous,<br />

randomly oriented polymer fibers are deposited on the collection plate.<br />

Electrospraying, the formation of droplets, or bead defects – fibers with beads on a<br />

string – can occur instead of continuous fibers depending on a large set of parameters. These<br />

include polymer properties such as molecular weight, solubility, and glass-transition<br />

temperature; solution properties such as concentration, viscosity, surface tension, electrical<br />

conductivity; and processing parameters such as applied voltage, plate distance from the<br />

spinneret, and solution flow. Even ambient properties such as relative humidity and solvent<br />

3


vapor pressure can have an effect on the resulting fiber morphology. Thus, it is demonstrated<br />

that although electrospinning appears to be a straight-forward process, it is actually<br />

multivariate problem that requires a high degree of optimization in order to achieve a<br />

successful result.<br />

Modification of the Electrospinning Setup<br />

Since the electrospinning setup is so simple, it is very easy to modify in order to<br />

change the structures and morphology of the resultant fibers. In the past, a rotating collector<br />

plate has been used to recreate both uniform and aligned fiber mats. 13 Arrays of multiple<br />

needles 14 have also been utilized in order to increase the productivity of the electrospinning<br />

process but potential problems with interference between needles also becomes an issue.<br />

Modification of the spinneret needle can also be performed to form a co-axial setup 15 which<br />

creates a single fiber with a core-sheath structure and enables the ability to electrospin<br />

immiscible polymer blends. A multi-axial 16 spinneret can also be constructed where multiple<br />

layers of a fiber can be created. In addition, by switching the typical placement of the anode<br />

and cathode in the electrospinning setup, it is possible to draw the more polarizable<br />

component to the surface. 17 Furthermore, the type of current, alternating or direct, has been<br />

found to have a slight impact on the degree of orientation and fiber density. 18 Needleless<br />

electrospinning has also been successful, and eliminates the burden of clogged needles, by<br />

utilizing magnetic fluids, 19 a pointed needle and the use of a tungsten electrode, 20 and<br />

conducting electrodes. 21<br />

4


The straightforward electrospinning setup discussed earlier creates randomly-oriented<br />

fiber mats, however, many applications, such as the fabrication of electronic or photonic<br />

devices, 22 will require well-aligned, unidirectional fiber constructs. Thus, the macroscopic<br />

alignment of electrospun fibers is one processing modification that has garnered a lot of<br />

attention. Another approach involves utilizing a rotating drum at high rotating speeds, which<br />

has been found to orient fibers along the winding direction 23 and also along the sharp edge of<br />

a tapered, wheel-like disc. 24 Both of these techniques have a high throughput, but<br />

electrostatic differences in attraction across the drum have led to different densities and<br />

alignment that is not perfect. Split electrodes are another common geometrical configuration<br />

utilized to produce aligned fibers. Additionally, uniaxial alignment can occur between two<br />

conductive strips separated by a variable gap (typically of several centimeters) on the<br />

collector plate as seen in Figure 1.3. 25 This technique allows for control over the aligned<br />

fiber density, length of alignment, and ease of removal for imaging or processing. Finally,<br />

four electrodes have also been used to form a cross-bar array of electrospun fibers, 26 a<br />

morphology needed for device fabrication. As electrospinning gets more developed towards<br />

scale-up and industrial applications, the ability to finely control the alignment of polymer<br />

fibers will become imperative.<br />

Other types of Electrospinning<br />

Traditional solutions used for electrospinning must have a sufficient viscosity in order<br />

to produce the appropriate surface tension and usually produces fibers with diameters less<br />

than 1 �m. For higher solution viscosities, filming can occur on the needle and impede the<br />

5


formation of fibers. An alternative is electrospinning from the melt,which often leads to<br />

much higher fiber diameters, but also must be performed under vacuum, and necessitates<br />

much larger electrode separations and thus electric fields. 27 Bowl electrospinning has a 40x<br />

increase in production rate and utilizes a bowl filled with a polymer solution and a concentric<br />

cylindrical collector. 28 Electrospinning polymers has also been proven to be effective using<br />

supercritical CO2 as the solvent 29 where solvent-free systems, specifically for biocompatible<br />

polymers, is needed. Another method involves blowing-assisted electrospinning, which is<br />

the use of a hot air stream for systems with very high viscosity. This has proven successful<br />

for hyaluronic acid, for example, a natural polysacchride with very high viscosity, in which<br />

no solvent was needed to create fibers. 30 Some downsides of this technique, however, are<br />

very large resultant fiber diameters, no increase in throughput, and a very large gas volume<br />

that is needed in addition to the expense of modifying the setup. 31 All of these methods have<br />

been formulated to adjust for limitations to traditional solution systems and maximize<br />

functionalities and target applications.<br />

1.1.2 Control over Fiber Characteristics<br />

External Structure<br />

Both the morphology and diameter of electrospun fibers depends on many of the<br />

processing and solution variables listed earlier. One of the more prevalent problems is the<br />

appearance of beads in the fibers, as seen in Figure 1.4. Of higher importance in the case of<br />

unidirectional fiber arrays is that any disturbance to the smooth fiber axis will negate the<br />

effect of alignment. Beads are often formed because surface tension drives the conversion of<br />

6


the polymer jet into spherical drops 1,32 as a way to decrease surface area. However, the<br />

forces acting against surface tension, specifically, electrostatic repulsion and viscoelastic<br />

forces, endeavor to form smooth fibers. Thus, in order to damper the appearance of beads, it<br />

is necessary to decrease the surface tension. This can be achieved by increasing the viscosity<br />

of polymer solutions, typically by an increase in molecular weight or concentration. The<br />

effect of this is clearly seen in Figure 1.5. Another way to eliminate beads is by the addition<br />

of salts 33 to increase the net charge density or blending of solvents 34 to lower the surface<br />

tension. Deitzel et al. have also determined that a threshold voltage exists 35 and when this is<br />

exceeded, it acts to weaken the stability of the jet and thus more beads are formed. Thus, as<br />

the above shows, there a multitude of ways to control bead density which is dependent on the<br />

specific polymer/solvent system.<br />

Fiber diameters in electrospun fibers are usually reported as a distribution and can be<br />

controlled by several means. 36 As more concentrated polymer solutions are used, fiber<br />

diameters become thicker. In addition, if the conductivity of a solution is increased, to<br />

reduce bead defects for example, then the resulting fiber diameter decreases. A higher feed<br />

rate will lead to thicker fibers. Rutledge et al. completed a comprehensive study on the<br />

multitude of variables that will affect fiber diameter and found that it is a fine balance<br />

between flow rate, the strength of the electric field, and the surface tension of the polymer<br />

solution that mostly affects fiber diameter. 37 Additionally, while fibers are typically circular<br />

in cross-section, other shapes can also occur. Ribbon-like structures with rectangular cross<br />

7


sections 38 have been found at higher viscosities and attributed to a film developing on the<br />

surface of the liquid jet and its subsequent collapse, as seen in Figure 1.6.<br />

Internal Structure<br />

One of the ways to impart functionality to fibers is by altering the internal structure.<br />

“Nanotubes” are electrospun fibers with a hollow interior that are most often created from a<br />

co-axial electrospinning setup and encompass the internal structure. This has applications in<br />

nanofluidics, drug delivery, and vascular engineering. Ceramic nanotubes are created by<br />

electrospinning two immiscible materials in a coaxial electrospinning setup in which the core<br />

can later be removed by solvent extraction. Ceramic nanotubes can be formed by the<br />

resultant calcination of the sheath at elevated temperatures. 39<br />

The benefits of using the coaxial spinning setup to form nanotubes is that the<br />

thickness of the sheath and inner diameter of the fibers can be varied by changing the<br />

spinning conditions 40 as seen in Figure 1.7. 39,41 Although these hollow fibers are very<br />

advantageous, they are also very fragile. In the past, this problem has been eliminated by the<br />

introduction an inorganic sol-gel precursor into the spinning solution and the subsequent<br />

formation of a gel network in the polymer sheath. 42 Nanotubules have also been created with<br />

porous surfaces, which were found to have higher rates of enzymatic reaction and are<br />

excellent candidates for applications requiring substrate penetration. 43 TiO2 nanotubes have<br />

been further functionalized by controlling the hydrophobic and hydrophilic properties on<br />

both the inner and outer surface and allowing adsorption of inorganic nanoparticles. 44 Fluid<br />

flow through these nanotubes would open up avenues for a variety of applications and flow<br />

8


inside carbonized hollow fibers has been demonstrated successfully with pressure-driven<br />

water in nanofluidic devices consisting of ~40000 nanotubes. 45<br />

In addition, electrospun fiber mats can be used as templates for the preparation of<br />

hollow fibers by the tubes by fiber templates (TUFT) process. 46,47 In this process the<br />

dissolvable electrospun fibers are coated with polymer, metals, or other materials. A<br />

negative replica is then created by the selective extraction of the template fiber. Composite<br />

fibers can then be created, for example, by coating poly(lactic acid)(PLA)/Pd(OAc)2 with<br />

poly(p-xylylene) (PPX) and pyrolyzing the PLA resulting with hollow PPX fibers with Pd<br />

nanoparticles on the interior 46 as seen in Figure 1.8. This production technique can be used<br />

to create ceramic nanotubes where a co-axial setup is not necessary, and has been performed<br />

with titania, 46 aluminum 48 and gold 49 just to name a few.<br />

Internal Poylmer Morphology<br />

A decade ago, the general goal of electrospinning was to establish which polymers<br />

could be electrospun from which solvents and gain a better understanding of how different<br />

solution/processing conditions affected the resultant morphologies. Now with this breadth of<br />

knowledge available, scientists are asking how to create a given morphology or property.<br />

Electrospinning is a fabrication technique that operates far from equilibrium. In fact,<br />

the time scale between a volume element leaving the jet to being collected is ~0.01<br />

seconds. 50 The evaporation of solvent and the elongation of the jet control both the internal<br />

and surface structure of the resulting fiber during this time frame and the choice of solvent<br />

can also affect the resulting fiber morphology, less volatile solvents may not completely<br />

9


evaporate and residual amounts could encourage chain relaxation. The elongation of the jet<br />

can result in crystalline polymers achieving chain orientation parallel to the fiber spinning<br />

direction, and crystal formation in electrospun fibers can even mirror those in processes with<br />

longer time scales, such as extrusion. 51 Similar to polymer thin films, annealing of fibers will<br />

convert crystals to a more ordered phase 52 but also requires the fiber to be encapsulated by a<br />

sheath material 53 to prevent flow.<br />

While electrospinning typically produces fibers with smooth surfaces, the creation of<br />

a porous surface resulting in an increase in surface area and is desirable for applications such<br />

as tissue engineering, filtration, and catalysis. Fiber mats can also be exposed to a solvent<br />

that induces swelling, once the swelling agent is removed then what remains is a fiber with a<br />

larger diameter and pores. One way porous fibers can be created is by the electrospinning of<br />

immiscible polymer blends. 54 For example, if a blended fiber is immersed in a selective<br />

solvent, one phase can be etched out and leave pores. Blended fiber mats can also be<br />

exposed to a solvent that induces swelling, once the swelling agent is removed then what<br />

remains is a fiber with a larger diameter and pores (Figure 1.9). Porosity can also be created<br />

by electrospinning fibers into a bath of liquid nitrogen, 55 which causes a phase separation of<br />

solvent and polymer.<br />

Although the structural characteristics of electrospun fibers are advantageous, the<br />

bulk properties tend to lack functionality that is desired for more multifunctional<br />

applications. One way of overcoming this problem is to create composite nanofibers,<br />

allowing the incorporation of two chemically and physically different components which can<br />

10


enhance the mechanical, 56 conductive, 57 and magnetic 58 properties, just to name a few.<br />

However, encapsulated molecules often show reduced activity 59 when constrained in a<br />

polymer matrix and may not always exist at the surface. For example, when antibacterial<br />

biocides are blended with a polymer prior to electrospinning their efficacy is limited and<br />

leave no potential to attack airborne pathogens. 60 In contrast, surface modification through<br />

the covalent bonding of poly(quarternary ammonium) can create a permanent antibacterial<br />

surface on a fiber. 61<br />

Because polymer surfaces exhibit typically low surface energy, 62 they need to be<br />

treated chemically or physically. Simple surface coating is usually the most straightforward,<br />

chemical modification for this treatment. In simple surface coating, electrostatic interactions<br />

or liquid-phase attachment are responsible for the deposition of conducting polymers or<br />

nanoparticles to the surface of electrospun fibers. 46 Physical or chemical vapor depositions<br />

can then be utilized to coat fiber mats with ceramics, polymers, or metals 48 as seen in Figure<br />

1.10. 63 In addition, precursor polymers can be spun, 64 such as poly(acrylonitrile)(PAN), 65<br />

and then subsequently carbonized to prepare ceramic fibers. Physical treatment techniques<br />

include plasma, 66 the formation of ‘layers,’ 67 ultraviolet treatment, 60 mineralization, 66<br />

etching, 68 or the inclusion of a composite material that is reactive. 69 Once there are<br />

chemically-active groups present on the surface, covalent bonding, 70 immobilization, 71 and<br />

electrostatic interactions 72 can be used to stabilize reactive groups to the surface of the fiber.<br />

Modification of just the surface of polymer fibers, not the bulk, can open these materials up<br />

11


for multifunctional applications where the fibers can interact with their environment such as<br />

filtering of target molecules, 73 protective textiles, 74 tissue scaffolds, 75 and drug delivery. 76<br />

Additional Properties of Electrospun Fibers<br />

Strength is an important consideration for many of the applications targeted by<br />

electrospun fiber scaffolds. Whereas tensile testing can easily be performed on macroscopic<br />

nonwoven fibers due to their larger size, the physical manipulation of nanofibers becomes an<br />

issue. One way to possibly alleviate this is by utilizing nano tensile testing on single fibers<br />

which has been reported for PLA. 77 While inconsistent results have been collected for tensile<br />

testing of a randomly oriented fiber mat, mechanical property trends have been found to exist<br />

for oriented fibers collected on a rotating drum as a function of the velocity of the drum. 78<br />

Nanoindentation is perhaps the best method to measure the elastic properties of a single<br />

fiber. 79 In situ tensile testing has been performed combining atomic force microscopy<br />

(AFM) and scanning electron microscopy (SEM) and it was determined that electrospun<br />

fibers had an increased tensile stress and elastic modulus as compared to bulk samples. 80 A<br />

uniform and thus comparative method of measuring the mechanical properties of both<br />

electrospun polymer fiber mats and individual fibers is a characterization technique that<br />

needs to be better established as this field progresses.<br />

1.1.3 Electropsinning Fiber Material<br />

One of the greatest benefits of electrospinning is the ability to easily tune the fiber,<br />

solvent, and processing parameters in order to create the desired material. Several classes of<br />

12


electrospinnable polymers will be briefly discussed as an overview and as a basis for the rest<br />

of this document.<br />

Types of Fiber-Forming Polymers<br />

Many of the applications for electrospinning are biological-based, ranging from tissue<br />

scaffolds 81 to wound dressing 82 to artificial blood vessels. 83 For this reason, the creation of<br />

electrospun mats utilizing biopolymers or blends of biopolymers is highly sought after and<br />

has broad application. Natural polymers tend to exhibit better biocompatibility when used in<br />

biomedical applications. Protein fibers, for example, including collagen, gelatin, elastin, and<br />

silk fibroin have been very well studied. 84 Although most of these materials generally lack<br />

strength, silk is unique in that there are multiple sources (silkworm 85 vs. spider) 86 and the<br />

prepared mats can be treated to induce conformational changes to beta-sheet structures to<br />

enhance its mechanical properties. 87 Polysacchrides such as dextran, 88 chitosan, 89 and<br />

cellulose acetate 34 have also been electrospun to form fiber scaffolds. In addition, many<br />

types of proteins and enzymes such as lipase, 90 cellulase, 91 and bovine serum albumin<br />

(BSA), 76 cannot be electrospun alone and thus are blended with biopolymers to create fibers.<br />

Fibers containing enzymes are often termed “bioactive” and, surprisingly, demonstrate<br />

increased enzymatic activity when immobilized in electrospun fibers. 76 Control over the<br />

release rate can be accomplished by either coating the external surface of the fiber or by<br />

coupling the enzymes to the fibers. 92<br />

Synthetic biopolymers, on the other hand, offer the advantage to tailor the resultant<br />

properties and are less expensive to fabricate. Hydrophobic polyesters such as poly(glycolic<br />

13


acid) (PGA), 93 PLA, 48 and poly(caprolactone) (PCL) 94 have all been easily electrospun.<br />

Hydrophilic biodegradable polymers such as polyurethane (PU), 95 poly(vinyl alcohol)<br />

(PVA), 36 poly(ethylene oxide) (PEO), 96 polydioxanone, 97 and polyphosphazene derivatives 98<br />

have been electrospun for biomedical applications. Synthetic copolymers have also been<br />

derived in order to combine desirable properties from two or more biopolymers such as<br />

mechanical robustness, morphology, pore size, biodegradability, cell affinity, etc. 99<br />

Poly(lactide-b-glycolide) (PLGA) is one copolymer that is especially well-studied and<br />

consists of a glycolide and lactide. Depending on the ratio of glycolide/lactide, the<br />

mechanical properties and biodegradability can be carefully tuned and have been tested for<br />

use as a antibiotic delivery vehicle. 100<br />

Water soluble polymers like PEO, PVA, poly(acrylic acid) (PAA),<br />

poly(vinylpyrrolidone) (PVP), hydroxypropylcellulose (HPC) are highly desirable for<br />

biomedical applications because the use of corrosive solvents can be avoided. PEO has been<br />

particularly well studied due to the range of molecular weights available, its solubility in a<br />

multitude of organic solvents in addition to water, 24,101 and its biocompatibility. The use of<br />

PVA is also advantageous since its degree of hydrolysis or water solubility can be varied. 102<br />

The reactive hydroxy groups also open up the possibility of reaction or modification. The<br />

solubility of polymers in water can also be adjusted by the pH value, temperature, or the<br />

addition of surfactants or solvents. 103 Although aqueous electrospinning eliminates the need<br />

for toxic solvents, applications such as filtration or use in textiles is not possible and,<br />

therefore, these mats are often crosslinked to improve water resistance. Although chemical<br />

14


crosslinking is very effective, 104 cross-linking agents can alter the properties of the polymer<br />

and lead to degradation. Thus thermal cross-linking has been investigated for blends of PVA<br />

and cylcodextrin 105 and photo cross-linking with PVA derivatives. 106<br />

There is a large list of organosoluble polymers as seen in Figure 1.11. 103 Organic<br />

solvents are useful for the ability to control volatility, conductivity, pH, and polarity. Many<br />

properties of organic solvents are not apt for industrial processes; namely flammability,<br />

toxicity, and corrosiveness. While there are many organosoluble polymers, just a few will be<br />

touched on in this chapter as we discuss the interesting properties that can result. First, PAN<br />

can be electrospun from dimethylformamide (DMF) and converted into carbon fibers via<br />

pyrolysis. 107 This creates a class of materials used as reinforcement in aerospace, industrial,<br />

and aerospace applications. 108 PAN fibers have also been reinforced with metallic oxide<br />

nanoparticles 109 and multi-walled carbon nanotubes. 110 Aliphatic poly(amide) (PA), or more<br />

commonly known as nylon, can be electrospun into very thin and uniform fibers but often<br />

require solvents that are very toxic. 111 Due to their high solvent and thermal stability, PA<br />

electrospun mats are also being investigated for filtration applications. 112 Poly(ethylene<br />

terephthalate) (PET), an aliphatic polyester, is known for its excellent structural and<br />

mechanical strength, transparency, and resistance to many solvents and has been electrospun<br />

in several applications. 113 The surface of PET can also be chemically modified to create a<br />

more functional surface. Poly(vinylidene fluoride) (PVDF), a polymer known for its piezo-<br />

electric properties, has been electrospun for applications in lithium ion polymer batteries. 114<br />

Finally, PU fibers are of interest due to their high flexibility and shape-memory properties 115<br />

15


and may have applications in wound dressings. This is simply a brief demonstration of the<br />

breadth of properties and applications that can be accessed through organosoluble polymers.<br />

Inorganic/Polymer Composite Fibers<br />

Although electrospinning is often associated with polymer fibers, there have been<br />

many reports of ceramic and inorganic fibers formed as well which have been touched upon<br />

already. These fibers can be created in three ways: electrospinning sol-gel precursor<br />

solutions, gas-solid reaction, or in situ photoreduction. The sol-gel method combines a<br />

colloidal solution (sol) which acts as a precursor to an integrated network (gel) of network<br />

polymers. 116 The main challenge of this technique is the creation of a viscosity that is<br />

equivalent to that of a viscous polymer solution. The hydrolysis rate of a sol-gel precursor<br />

can be controlled by changing the pH value or aging conditions. 103 The direct<br />

electorspinning of viscous inorganic sols have successfully created fiber mats of<br />

TiO2/SiO2, 117 PbZrxTi1-xO3, 118 and SiO2 119 fibers. A slightly different approach to the<br />

formation of inorganic fibers is through the use of a series of mesoporous molecular sieves<br />

for the electrospinning of sols and structure-directing agents. 120 This approach involves the<br />

formation of ceramic fibers when the organic portion is removed by calcination at elevated<br />

temperatures and has been applied to anatase fibers 25 and many other metal oxides. Non-<br />

oxide ceramic nanofibers have also been prepared by electrospinning a sol solution<br />

containing Novolac resin and tetraethylorthosilicate followed by conversion to SiC through<br />

pyrolysis. 117 Electrospinning of inorganic and ceramic fibers may become more widespread<br />

16


especially when considering applications requiring a strong structure, membranes, catalytic<br />

supports, or actuators.<br />

Gas-solid reactions were initially introduced to incorporate semiconductor<br />

nanostructures into electrospun fibers. 121 This technique involves co-dissolving a metal salt<br />

and polymer into one solvent and then electrospinning to obtain a polymer/metal salt<br />

composite nanofiber which can then be exposed to a H2S gas at room temperature to<br />

synthesize PbS nanoparticles in situ. The color of the composite quickly turns from white to<br />

yellow after exposure to the gas. 122 Both 20 nm diameter spherical nanoparticles 123 and<br />

nanorods have been obtained on the surface and internally through this technique (Figure<br />

1.12). Chloride nanoparticles have also been obtained from exposure to hydrochloric acid<br />

(HCL). 121 Although this technique consists of multiple steps, it is possible to create<br />

nanofibers with metallic nanoparticles through the simple exposure to a gas.<br />

Metal colloids can also be created through photochemical reduction. UV irradiation<br />

of aqueous solutions of AgNO3 induces photo-oxidation of water and results in the formation<br />

of silver atoms. 124 This same idea can be applied to polymer solutions where the composite<br />

fibers are UV treated and have been demonstrated with CA/AgNO3 composite fibers 125<br />

which resulted in the formation of silver nanoparticles as small as 5 nm were formed. Gold<br />

colloids from HAuCl4 can be reduced under UV irradiation but require an organic stabilizer<br />

such as PVP or PVA. 63 In addition, a photocatalyst of TiO2 is also required and these<br />

applications could include chemical or biological sensing. Although in situ formation of<br />

17


metallic nanomaterials can result in high and uniform loadings, they also usually require the<br />

use of catalysis and external modifications in order to be successful.<br />

1.1.4 Imparting Functionality to Electropsun Fibers<br />

Functional, organic fibers can be created by combining two or more polymers with<br />

different characteristics; this can encompass polymer blends, block copolymers, and<br />

encapsulating functional materials into the electrospinning solution.<br />

Multicomponent Polymer Systems<br />

Both miscible and immiscible polymer blends can be electrospun in order to create<br />

new morphologies of phase separated nanofibers. When two polymers phase separate into<br />

discrete domains, one component can be selectively removed in order to form porous fibers.<br />

This has been conducted with PLA/PVP, 54 PEO/silk, 126 and PGA/chitin, 127 When more<br />

controlled phase separation occurs, core-shell structures 128 or semiconducting wires 129 can be<br />

formed. A core-sheath structure is highly desirable due to the fact that it is possible to<br />

combine polymers with two different sets of properties; i.e. the sheath can be insulating while<br />

the core can be conductive. 128 In addition, the core can also be etched to form nanotubes 39<br />

which may be beneficial for drug release 130 and sensors. 131 Core-sheath structures (Figure<br />

1.13a) can be formed via careful solvent selection, 132 due to thermostatic/kinetic differences<br />

between two polymers, 133 utilizing a co-axial spinning method, 134 or induced by an electric<br />

field. 17 When melt polymer blends are electrospun, it is also possible for a chemical reaction<br />

to occur. With the broad range of polymers and their respective properties, there are endless<br />

combinations of blend fibers that can be created by varying the ratio, molecular weight, and<br />

18


number of components. Molecular weight control of one component can be used to tune the<br />

resulting properties of the fiber mat, for instance by changing the molecular weight of PVA<br />

in a blend system it was possible to reduce the formation of beads. 135 This can be a favorable<br />

process, as in the case of copolymerization reactions during melt electrospinning. 136 Some<br />

benefits to using polymer blends to create nanostructured fibers are careful control over<br />

microphase size based on both molecular weight and polymer:polymer ratio, a direct control<br />

over hydrophobicity/phillicity based on polymer choice, and ease of setup.<br />

Electrospun block copolymer systems have restricted phase separation and usually<br />

micro-phase separate into domains with sizes less than 100 nm. 137 Poly(styrene)-b-<br />

poly(isoprene) (PS-b-PI) has been found to phase separate in electrospun fibers into both a<br />

cylindrical and lamellar morphology parallel to the fiber surface as shown in Figure 1.13b-<br />

c. 138 A block copolymer/amphiphile system, PS-b-poly(2-vinyl pyridine) (P2VP)/<br />

poly(diallyl phthalate), was also found to phase separate into elongated spherical domains in<br />

thin films and the bulk. Fibers with controlled hydrophobicity or hydrophilicity can also be<br />

achieved by electrospinning with block copolymers. 139 Electrospun block copolymer<br />

systems are most applicable for biomedical applications where the decomposition and<br />

biocompatibility rate of the block copolymer can be carefully tuned. 140,141 Although the<br />

ability to harness the periodic micro-phases that can be created with an electrospun block<br />

copolymer system is highly appealing, a significant challenge is the aspect of annealing. In<br />

thin or bulk films, the shape of the system is not impacted when annealing occurs. With a<br />

fiber however, flow can cause a loss of shape. This loss of shape has been avoided by Kalra<br />

19


et al. where the fibers are coated in silicon prior to annealing. 53 The question that remains is<br />

whether this process has scale-up capability but whether block copolymers remain a great<br />

way to create multifunctional polymer nanofibers.<br />

Encapsulation of Functional Materials<br />

Incorporation of inorganic nanomaterials into polymer matrices can yield hybrid<br />

polymer nanocomposites with synergistic properties. These materials are often added to the<br />

polymer solution prior to electrospinning but can also be formed in vivo from precursors in<br />

the polymer solution.<br />

Carbon nanotubes (CNTs) are of particular interest to scientists due to properties such<br />

as electrical, electronic, and thermal conductivity 142 as well as its high strength, flexibility,<br />

and resilience. 143 These properties can be imparted to the nanofiber when CNTs are<br />

combined with the polymer solution prior to electrospinning. It has been well documented<br />

that CNTs orient with their long axis parallel to the electrospun fiber axis, as seen in Figure<br />

1.14. 144 Both single-walled 145 and multi-walled 146 have been electrospun but often require<br />

high concentrations 147 or use of a surfactant in order to get good dispersion.<br />

Functionalization of CNTs can be accomplished through esterification 148 or oxidation 110 and<br />

directly affects the mechanical properties, where functionalized CNT had twice the tensile<br />

strength of unmodified CNTs. 148 This was most likely due to better dispersion and better<br />

interaction between the nanotube and the polymer matrix. A theoretical model was<br />

formulated by Cohen et al. and it was determined that CNTs align along the jet stream<br />

immediately prior to being expelled from the spinneret due to the sink-like flow leading to<br />

20


the jet. 149 The electrical conductivity of CNT polymer composites was also tested and found<br />

to reach 10 -2 S cm -1 . 150 Indeed, the use of CNTs continues to be investigated as a reinforcing<br />

agent.<br />

Although many applications would benefit from encapsulated enzymes or other<br />

bioactive species within a fiber matrix, a large challenge exists in maintaining their activity<br />

post-electrospinning. They are usually sensitive to heat and can lose activity when exposed<br />

to solvents or other chemicals. Bioactive agents such as hyaluronic acid, 151<br />

glycosaminoglycan, 152 heparin, 153 and bone morphogenic proteins 154 have all been<br />

incorporated into polymer fiber matrices without losing their activity. One target application<br />

of these composite nanofibers is for the detection of molecules such as ammonia 155 or<br />

glucose. 156<br />

Other notable, functional materials that have been electrospun include nanoplatelets,<br />

carbon black, graphene, and quantum dots. The most common nanoplatelet is<br />

montmorillonite (MMT) and its purpose is usually that of a reinforcement in polymer<br />

matrices to improve mechanical strength and thermal stability. 157 Fiber size plays an<br />

important role in this, as Li et al. determined that when smaller Nylon/MMT 158 fibers had a<br />

tensile strength than composite, fibers four times the size which was attributed to the<br />

crystallinity of the nylon (Figure 1.15). Carbon black has also been encapsulated into<br />

electrospun fibers to increase either electrical conductivity or the nanofiber modulus, 159 and<br />

to sense strain. 160 Nanoparticles such as calcium carbonate are also being used for<br />

biomedical applications such as bone regeneration. 161 The direct dispersion of graphene<br />

21


prior to electrospinning has been used to enhance the optical absorption of poly(vinyl<br />

acetate) (PVAc) by a factor of ten. 162 Quantum dots (QDs) are also of interest due to their<br />

interesting optical properties because of a quantum confinement effect. In fact, the high-<br />

voltage source in electrospinning can actually effect the passivation of the QD and suppress<br />

deep-level emission. If the voltage is increased high enough, it may be able to align the ZnO<br />

QDs and lead to ultraviolet photoluminescence. 163 Aggregation is a problem with these<br />

systems and the addition of surfactant has been shown to allow a uniform distribution for<br />

CdTe QDs in a PVP matrix. 164 Cellulose nanocrystals have also been electrospun with a<br />

PVA nanofibers and were found to significantly increase the elastic modulus while acting to<br />

reinforce the nanofibers. 165 Additionally, superhydrophobicity has been created by<br />

electorspinning epoxy-siloxane modified SiO2 nanoparticles with a diameter of 200 nm in a<br />

PVDF matrix. 166 Metal nanoparticles have also been used as localized heat sources inside<br />

PEO fibers when irradiated with a laser tuned to the surface plasmon resonance of the noble<br />

metal nanoparticle and complete melting was observed. 167<br />

Although it has been determined that nanoparticles located in the interior of a<br />

polymer fiber can show reduced activity, there is also great concern about the toxicity and<br />

release of the nanoparticles into the environment. For this reason, the goal of many scientists<br />

is to increase the functionality of a nanofiber mat without toxicity, thus dispersing<br />

nanoparticles on the interior of the fiber surface has been achieved. Electrospun fibers with<br />

noble metal nanoparticles such as gold, 168 silver, 169 and manganese acetate 170 have been<br />

produced to attempt to achieve this result. Silver nanoparticles can impart antimicrobial<br />

22


properties to polymer fibers 171 since they are positively charged and attract electronegative<br />

bacteria. Gold nanomaterials, on the other hand, exhibit strong surface plasmon resonances<br />

in the visible electromagnetic spectrum as a consequence of optically-driven coherent<br />

oscillations of conduction electrons. 172 These resonances give rise to characteristic optical<br />

absorption and scattering spectra, 173 yielding brightly colored, nanoparticle-containing<br />

suspensions (Figure 1.16). Palladium nanoparticles have specific catalytic activity in<br />

hydrogenation of dienes and olefins, and has been electrospun with PAN. 174 Magnetic<br />

particles such as iron oxide have also been blended in electrospun polymer fibers 175 and have<br />

also been found to orient parallel to the fiber axis, deflect under a magnetic field, and<br />

improve the mechanical properties of the fibers. The formation of nanocomposites can even<br />

change the inherent properties of the filler material; core-shell Fe/FeO nanoparticles have<br />

been found to increase in shell thickness by a factor of 7.4% when electrospun. 176<br />

1.1.5 Applications<br />

The main attributes of electropun fiber webs are: high surface area to volume ratio,<br />

porosity, and low mass. Many applications exist based on the ability to functionalize the<br />

nanofibers as discussed earlier in this chapter. The types of applications and the required<br />

fiber characteristics will be discussed in this section.<br />

Tissue Engineering<br />

Electrospun fiber mats are applicable for regenerative tissue due to the similarity to<br />

the extracellular matrix (ECM), which provides the structural support to the cells in tissue<br />

and organs. 177,178 Biocompatabile and biodegradable polymers can be used to make an<br />

23


artifical ECM and seed stem or human cells onto it. When comparing cell proliferation on<br />

both fiber and film of the same polymer, the fibers had a positive effect on cell growth. 179<br />

Some challenges that exist, however, include the creation of a smooth fiber surface for cell<br />

growth, 178 which can be created by altering the processing parameters. Porosity is another<br />

important challenge, as it can cause the cells to create bridges over the pores in a fiber<br />

scaffold. 180 Different tissues have drastically different tensile strengths, elastic moduli, and<br />

elongational strength (i.e. skin vs. cartilage) and this has to be taken into consideration when<br />

choosing a polymer. The polymer choice must allow the anchorage, migration, and<br />

proliferation of cells in addition to matching the mechanical properties of the target tissue<br />

and include many of the natural and synthetic biopolymers listed earlier.<br />

Polymer blends are also utilized in order to combine properties; i.e. biodegradability<br />

with mechanical robustness. 181 Additives such as calcium carbonate, phosphate, and<br />

hydroxyapatite can increase functionality for tissue such as bone. 182 The types of cells that<br />

have been seeded include mesenchymal stem cells, endothelial cells, neural stem cells,<br />

fibroblasts, osteoblasts, etc. 103 One challenge that still remains is the creation of a synthetic<br />

ECM as a hierarchical complex structure, to be fully functional the structure and growth of<br />

arteries, 183 etc. may also need to be taken into account. The physical properties of the mat,<br />

such as using aligned fibers, has been shown to guide the growth and orientation of<br />

neurons. 184 A three-dimensional assembly of aligned, mineralized fibers 185 has been created<br />

by Teo, et al. to mimic the hierarchical structure of bone and an image of these bundles can<br />

be seen in Figure 1.17.<br />

24


Catalysis<br />

An important step in catalysis is the removal of catalyst after the reaction, to avoid<br />

contamination in the end product. If catalyst are encapsulated within a nanofiber matrix, this<br />

could be alleviate that problem by either dipping the nanofiber mat into the reaction solution<br />

or have the solution pass through it in a reactor. 186 Polymers can be electrospun with<br />

encapsulated monometallic or bimetallic nanoparticles such as Rh, Pt, Pd, Pd/Pt, or Rh/Pd for<br />

catalysis in hydrogenation reactions. 174 In contrast, nanoparticles can be formed in vivo by<br />

electrospinning solutions containing metal salts and reducing these agents at high<br />

temperatures. Core-shell fibers created by the TUFT method for homogeneous catalysis are<br />

able to achieve conversion in shorter times relative to conventional catalysts and can be used<br />

several times without a loss of activity. 187 Carbon fibers decorated with Pt nanoparticles<br />

were fabricated for electrocatalysis and had excellent activity and stability towards the<br />

oxidation of methanol. 188 This demonstrated that carbon fibers loaded with noble metal<br />

catalysts are advantageous due to the high conductivity and close contact between carbon<br />

matrix/Pt particle, careful control over particle dispersion would be necessary to enhance this<br />

application.<br />

Drug Release<br />

Electrospun fibers for drug delivery are advantageous for the ability to control the<br />

release rate, easy production and polymers have little influence on the carrier. 189 In order to<br />

be successful in this application, nanofiber mats need to fulfill several requirements. First,<br />

they need to possess some robustness in order to prevent the drug from dissolving before it<br />

25


eaches its target and allow for a controlled release if possible. The drug release could be<br />

triggered by a stimulus and complete its release by a certain point. Nanoparticles of lipids<br />

and biopolymers have been investigated for the purposes of drug transport and release. 190<br />

Nanomaterials are also be used for locoregional therapy for immediate release to a target area<br />

and is applicable for wound healing or tissue engineering. Nanofibers loaded with iron oxide<br />

nanoparticles can be subjected to an external magnetic field which in turn causes heating of<br />

the nanoparticles. This heat can be the stimuli needed for drug release. 76,191 A controlled<br />

release is essential for many therapies and this is a functionality that many traditional<br />

nanofbiers lack. By utilizing core-shell fibers the core can immobilize the drugs while the<br />

shell controls the release rate out of the fibers. 131 Core-shell fibers can also be loaded with<br />

proteins, enzymes, and growth factors for a controlled release. 141,192<br />

Water & Air Filtration<br />

Electrospun fiber mats are currently being investigated as inexpensive, low-energy<br />

filtration membranes in order to provide clean water. The high volume to surface area and<br />

nanosize pores make them ideal to remove contaminants like particles, organisms, and<br />

hazardous chemicals from water. While the mechanical stability can be tuned with the<br />

choice of polymer, the question still remains whether it can withstand a high number of water<br />

cycles and still be effective. Recent reports of an electrospun nanofiber mat being used for<br />

microfiltration have demonstrated that 99% of all particles that passed through were collected<br />

and no failure was observed. 68,193 For the removal of heavy metals, the addition of heavy<br />

metals on the surface, for example, NH3+ functional groups, 194 carboxylic acid, 195<br />

26


poly(methacrylic acid), 73 has a higher success rate than encapsulated materials due to a<br />

reduced activity. These membranes can also be imparted with antimicrobial properties<br />

through the incorporation of silver nanoparticles, 196 sulfonated phenol groups, 197 and n-<br />

halamine. 198 Water filters consisting of boehmite nanoparticles impregnated in electrospun<br />

fibers was found to have an excellent removal efficiency of cadmium (II) from water. 199<br />

Functionalizing the surface of fiber mats may be as simple as exposure to a basic solution;<br />

antimicrobial properties have been found for PAN fibers with amidoxime groups when<br />

exposed to a NH2OH solution. 200<br />

Air filtration covers everything from dust to viruses to warfare stimulants. Nanofiber<br />

membranes are already in use in air filtration media by several companies 201 based on their<br />

high efficiency, ability to trap smaller particulates, and large surface area. Smaller fibers do<br />

have a larger pressure drop when compared with macroscale fibers, which is one downside.<br />

Much like water filtration, air filters can be chemically functionalized for detoxification<br />

purposes against neurotoxins. 202 Likewise, ceramic fibers such as zinc titanate had an<br />

excellent decomposition efficacy against nerve agents. 203 The future of filtering technology<br />

looks to be in functional, antimicrobial and detoxifying membranes but it will be necessary to<br />

absolutely determine that none of the incorporated additives leach out of the nanofibers<br />

overtime.<br />

Sensors<br />

Detecting contaminants in the air can be a key feature in environmental strategy.<br />

There are many air pollutants that are both low in concentration and low in size and<br />

27


nanofiber template sensors would be greatly applicable for both of these characteristics. For<br />

liquid applications, trace amounts of contaminants such as pharmaceuticals, cosmetics,<br />

hormones, etc are possible carcinogens and can cause health problems. 204 Molecularly<br />

imprinted nanoparticles in polymer fibers can detect very small amounts of propranolol in tap<br />

water, 113 even though they are not on the surface. This same idea could be applied to<br />

detecting other pharmaceuticals in waste water. Nanofiber mats can also be used as gas<br />

sensors in areas with high air pollution. These are often ceramic materials, molybdic oxide<br />

and tungsten oxide fibers have been shown to be successful in detecting NO2 and NH3 in<br />

air. 155 Nanofiber sensors are superior to other methodologies due to their fast response time<br />

and detection limit sensitivity.<br />

1.2 One-Dimensional Magnetic Nanostructures<br />

As technology advances, electronic devices are getting both smaller and lighter. In<br />

order to realize the concept of nanodevices, a replacement has to be found for electrical<br />

circuits and electrical wiring. One candidate for nanowires is actually deoxyribonucleic acid<br />

(DNA); it has excellent self-assembly and has the potential for nanoscale patterning.<br />

Conductivity can be imparted by replacing the imino proton on each base pair with a Zn 2+<br />

ion. 205 Ferromagnetism has also been incorporated when DNA is in the presence of Cu 2+<br />

ions, the spacing between the base pairs mimicks a ferromagnetic material. 206 Any polymer<br />

with a saturated backbone of pendant metal complex has the potential to be of use in memory<br />

applications.<br />

28


One-dimensional nanostructures have very unique properties that act on two dimensions;<br />

properties of the individual nanoparticles but also anisotropic properties that exist<br />

collectively. A multitude of properties can exist when the unidirectional arrangements of<br />

these materials exist within an organic matrix. The presence of an organic matrix helps these<br />

materials, which usually tend to self-aggregate, keep their shape and approve their stability.<br />

These are of interest for sensing, magnetomechanical systems, spintronic devices, and data<br />

storage systems. When an external magnetic field is applied to metal-containing atoms such<br />

as Co, Ni, and Fe, the spins of unpaired electrons become aligned. The term ‘magnetism’<br />

refers to how materials act on a molecular level to an external magnetic field. The magnetic<br />

susceptibility, �, is the effectiveness of an applied field to induce a magnetic dipole. It is a<br />

ratio of the induced magnetization, M and the applied field: � = M/H. The behavior of<br />

materials in a magnetic field can be broken up into the following: 207<br />

A) Diamagnetism: This is a basic property of all materials which are weakly repulsed by<br />

a magnetic field.<br />

B) Paramagnetism: There are unpaired electrons in this material whose magnetic<br />

moments will align in the same direction as the applied field.<br />

C) Ferromagnetism: The electron spins are coupled into a parallel alignment that is<br />

maintained by thousands of atoms in magnetic domains. The number of domains<br />

depends on the size of the material; materials below a critical diameter of 14nm<br />

consist of just a single domain. Above this critical diameter, the material will show<br />

remanence or a persistent magnetization where some directional domains still exist<br />

29


even when the magnetic field is changed and coercivity, or the need for a magnetic<br />

field in the opposite direction to demagnetize the material. The bulk property for<br />

larger materials is that many magnetic domains where the spins point in the same<br />

direction and act cooperatively.<br />

D) Antiferromagnetism: In this type of substance the electron spins of atoms are fixed in<br />

an antiparallel alignment. These materials have a zero net magnetic moment.<br />

E) Ferrimagnetism: A strong net dipole is present due to a range of spin moments<br />

aligned in an antiparallel direction. The critical diameter of a single domain of<br />

magnetite, a ferromagnetic material, is 128 nm.<br />

F) Superparamagnetism: This occurs when the size of a ferro- or ferrimagnetic particle<br />

is below the critical diameter and a stable magnetization is not possible due to<br />

Brownian rotation, or thermal fluctuations, of the material. The particles mimic that<br />

of a paramagnetic spin with a much higher moment. A critical temperature, called the<br />

blocking temperature Tb, exists where the dipole moments are able to align and<br />

couple to form a larger collective magnetization. Above this temperature, thermal<br />

fluctuations are the dominant force. These materials usually do not show any<br />

coercivity or remanence magnetization at room temperature.<br />

The reason scientists want to align magnetic materials is due to the shape anisotropic<br />

properties that exist which can lead to some interesting magnetic properties. Crystal<br />

anisotropy is due to the shape of the material; it is easier to magnetize rods or cylinder-<br />

shaped nanomaterials along the long axis as opposed to the short one. Some of the ways<br />

30


magnetic unidirectional materials are being fabricated include a template-based approach, 208<br />

channels in solids, 209 template through block copolymer micro-phase separation, 210<br />

cylindrical polymer brushes, 211 biological 1-D templates such as bacterial chains, 212 (seen in<br />

Figure 1.18) and electrospinning. 213 It is conducive to utilize an organometallic matrix<br />

because even very small concentrations of magnetic materials, i.e. less than 1%, can impart<br />

magnetic properties to the material as a whole. Multi-segmented alloys can even act as tiny<br />

magnets in solution by magnetizing only one of its components, as in the case of nanowires,<br />

and leading to self-assembled nanowires in an unidirectional arrangement. 214<br />

Superparamagnetic nanoparticles have been studied as possible contrast agents in MRI.<br />

Nanomaterials would offer improved sensitivity over the currently used contrast agents;<br />

chains of iron oxide nanoparticles within a biopolymer layer have the ability to circulate,<br />

target, and image tumors. 215 Although toxicity in the human body is a clear concern, it has<br />

been demonstrated in rats that these nanomaterials can be quickly expelled from the brain<br />

area into the circulatory system. 215,216 Harnessing spatial positioning over unidirectional<br />

magnetic nanoparticles can lead to multifunctional materials in a variety of applications.<br />

1.3 Organometallic Polymers<br />

1.3.1 Overview<br />

The first chapter focused on different ways to impart functionality to polymer fibers<br />

and some of the drawbacks of these approaches while the second chapter investigated the<br />

benefits of utilizing unidirectional magnetic materials. What if instead of having a<br />

multicomponent system, it was possible to have a metal in the polymer center which still<br />

31


imparted interesting properties while maintaining conventional polymer processing?<br />

Organometallic polymers make this possible, these polymer contain a transition metal in the<br />

main chain or more specifically has a metal-carbon σ or π bond. 217 The term ‘transition<br />

metal’ refers to any element that has an incomplete d sub-shell or which can give rise to<br />

cations with an incomplete d sub-shell. 218 The properties that make these transition metals<br />

interesting, however is the ability to have multiple oxidation states which can be easily<br />

controlled through electric currents. The first organometallic polymer was synthesized in<br />

1955 as poly(vinylferrocene) and displayed reversible oxidation-reduction properties. 219 The<br />

incorporation of metallic elements into polymer systems would allow different coordination<br />

numbers and geometries and thus supply fascinating magnetic, optical, or catalytic<br />

properties. Structurally, metallopolymers can be broken down into three categories: metals<br />

incorporated directly into the polymer chain, π or σ-coordinated metals, and metallic moieties<br />

pendant to the polymer backbone or in the side chains. 220 Some challenges that have existed<br />

in the synthesis of organometallics include low molecular weights, oligomers, impurities, and<br />

insolubility.<br />

Other high-yield organometallic polymers are often synthesized through a step-<br />

growth polymerization mechanism to create coordination chain polymers. Coordination<br />

polyelectrolytes have been synthesized through the self-assembly of bisterpyridines and<br />

metal ions where the polymers are water soluble and whose optical properties scan the visible<br />

spectrum. 221 The first bimetallic polymer was synthesized containing both ruthenium and<br />

silver and contains a ligand as a spacer. 222 Although organometallic polymers have been<br />

32


synthesized through this coordination step-growth polymerization, there is often little control<br />

over the final molecular weight or polydispersity. A more precise approach ca be achieved<br />

through coupling two polymer blocks through a living organic polymerization and metal<br />

coordination interactions. 223 A multitude of block copolymers can be created with a single<br />

metal atom joining the two blocks as reported by Shubert et al. 224 Though this synthetic<br />

method, one of the block components can be etched out leaving a regularly-ordered metal-<br />

polymer block copolymer thin film as seen in Figure 5.19. It is even possible to create<br />

metallogels, utilizing ligands and silver salts it is possible to form cages of coordination<br />

complexes. Other favorable properties possible from metallogels include fluorescence<br />

enhancement, 225 catalysis, 226 templating, 227 and electrical actuators. 228 The incorporation of a<br />

metal atom into the main chain of a polymer leads to traditional polymeric properties while<br />

maintaining the functionalities of metals including redox, magnetic, catalytic, and optical<br />

activity.<br />

1.3.2 Poly(ferrocenylsilanes)<br />

In the early 1990’s, an organometallic polymer was fabricated by Manners et al. and<br />

this was poly(ferrocenylsilanes) (PFS) with a high molecular weight and narrow<br />

polydispersity through a ring-opening polymerization method 229 and this synthesis technique<br />

has since yielded other strained monomers including bridging elements such as germanium,<br />

tin, and phosphous. 230 PFS can be tuned to have either a semi-crystalline or amorphous<br />

configuration due to the constituent groups on the silicon atom 231 where symmetrically<br />

substituted constituents, R=R’=Me, impart crystallinity. The prevalence of iron in the main<br />

33


chain conveys some interesting properties not present in non-metal containing polymers such<br />

as redox-activity due to the reversible (Fe(II)/Fe(III) couple) 218 , the ability to be pyrolyzed<br />

into a magnetic ceramic, 232 and semi- and photo-conductivity. 233 Gelable PFS derivatives<br />

have been electrospun and crosslinked, and strain-induced buckling on electroactuation was<br />

found to occur. 234<br />

Block copolymers are macromolecules containing two or more types of repeat units<br />

within long contiguous sequences, or “blocks,” of the same unit. These sequences are<br />

covalently linked to form a single macromolecule. They can spontaneously self-organize or<br />

microphase separate into a variety of ordered nanoscale morphologies such as lamellae,<br />

hexagonally packed cylinders, spherical micelles on a body-centered-cubic lattice, or<br />

complex bicontinuous morphologies like the gyroid. The type of morphology exhibited by<br />

the block copolymer depends on the chemical attributes and lengths of the blocks, the<br />

molecule’s architecture, temperature and overall chain length as well as presence of an<br />

additive like a solvent, homopolymer or another block copolymer. 235 Due to their ability to<br />

microphase separate, block copolymers constitute a versatile platform for a number of<br />

existing and emerging technologies such as adhesives, membranes, drug delivery,<br />

biomaterials and lithography. 236<br />

The formation of metal containing block copolymers was possible due to the controllability<br />

of the anionic ring opening polymerization of silicon-bridged ferrocenes. 237 Polymerization<br />

must occur through sequential addition of monomers with decreasing end-group reactivity<br />

such that PS ~ PI > PFS > poly(dimethylsiloxane) (PDMS) 237 . The first two PFS-containing<br />

34


lock copolymers were PS-b-PFS and PFS-b-PDMS 238 and have allowed the synthesis of<br />

block copolymers of PFS and PI, 239 poly(methylmetacrylate) (PMMA), 240 P2VP,<br />

poly(ferrocenylethylmethylsilane) (PFEMS), 241 and the hybridization with polypeptides. 242<br />

PFS block copolymers have been shown to self-assemble in the solid state and have<br />

demonstrated spherical, cylindrical, lamellae, and gyroid morphologies. 243 The lengths of<br />

poly(ferrodimethylsilane) PFDMS block copolymer micelles can elongate through the<br />

addition of unimers which act to couple existing micelles. 244 Triblock copolymers can be<br />

selectively etched so that only PFS is remaining as seen in Figure 3 which gives rise to many<br />

lithographic applications. In the solid state, amorphous PFS with unsymmetrical constituents<br />

on the silicon atom is utilized in order to allow phase separation to occur. Co-micelles with a<br />

scarf-shaped architecture (Figure 1.20) have likewise been prepared with platelet micelles as<br />

intiators. When a second PFS block copolymer is added, cylindrical micelle tassles are<br />

grown from the already present micelles. 245 Segregated micelles can also be used as a<br />

template for the formation of metallic particle and has been achieved with gold nanoparticles,<br />

PbS quantum dots, and titania. 246 Banded light-emitting barcode structures with fluorescent<br />

segments of a triblock-containing PFS were synthesized separated by nonemissive segments<br />

of a diblock. 247<br />

The periodic PFS domains can be converted to iron-rich clusters within a ceramic<br />

domain by pyrolysis at temperatures about 600°C which can lead to the growth of single-<br />

walled carbon nanotubes for soft lithography, 248 the formation of magnetic ceramics, 249 and<br />

PFS-derived catalysts. 250 Solution self-assembly of diblock copolymers occurs when a block<br />

35


selective solvent is utilized in order to induce segregation of the solvent incompatible ‘core’<br />

surrounded by the solvent compatible ‘corona.’ Micelles of PFS-containing block<br />

copolymers have produced micellar morphologies such as cylinders, 251,252 tubes, 252 fibers, 253<br />

and tapes. 239 PFS experiments in the bulk utilize PFDMS, whose crystalline nature is<br />

responsible for the growth of micelles. 254 Initial experiments with PFDMS-b-PDMS (with a<br />

block ratio of 6:1 respectively) in hexanes demonstrated that cylinders with a PFS core and<br />

PDMS corona were formed. 251 Imaging these organometallic, self-assembled structures with<br />

transmission electron microscopy (TEM) is easy due to the contrast already inherent in the<br />

sample owing to the iron-rich core. In order to form a core of PFDMS in cylindrical micelles<br />

it is necessary to have a corona:core ratio of at least 5:1 239,251 but when this ratio reaches 12:1<br />

it is more common for hollow, tube-like structures to form. 240,252 The length of the<br />

cylindrical micelles can also be easily controlled; the addition of a small amount of a<br />

common solvent will allow growth of the micelles and ultrasound or high temperatures will<br />

cleave longer cylindrical micelles. The self-assembly of PFDMS-b-P2VP was investigated<br />

in several alcoholic solvents and it was noted that cylindrical micelles were formed in<br />

isopropanol but not ethanol. 255 The reason for this was attributed to the fact that isopropanol<br />

is a better solvent than ethanol based on solubility parameters for PFDMS, which gives the<br />

PFS chains time to rearrange and crystallize more easily. PFS-b-PI formed cylindrical<br />

micelles in a PI selective solvent and the vinyl groups in the PI corona were utilized to<br />

conduct a Pt(0)-catalyzed shell crosslinking reaction which was a precursor to making PFS<br />

nanoceramics by pyrolysis with shape retention. 256 Nanocomposite self-assembled structures<br />

36


consisting of PFS-b-poly(vinylmethylsiloxane) (PVMS) wormlike micelles reacted with<br />

Ag[PF6] to create a one-dimensional array of silver nanoparticles encapsulated within the<br />

worm-like micelles 257 allowing for highly oriented arrays of nanoparticles to be formed.<br />

37


Nomenclature<br />

AFM Atomic force microscopy<br />

BSA Bovine Serum Albumin<br />

CNT Carbon nanotube<br />

DMF Dimethylformamide<br />

DNA Deoxyribonucleic acid<br />

ECM Extracellular matrix<br />

HCL Hydrochloric acid<br />

HPC Hydroxypropyl cellulose<br />

MMT Montmoillonite<br />

PA Poly(amide)<br />

PAA Poly(acrylic acid)<br />

PAN Poly(acrylonitrile)<br />

PCL Poly(caprolactone)<br />

PDMS Poly(dimethoxysilane)<br />

PFDMS Poly(ferrodimethylsilane)<br />

PFEMS Poly(ferroethylmethylsilane)<br />

PFS Poly(ferrocenylsilane)<br />

PGA Poly(glycolide)<br />

PI Poly(isoprene)<br />

PLA Poly(lactic acid)<br />

PLGA Poly(lactide)-co-(glycolide)<br />

PMMA Poly(methyl methacrylate)<br />

PPX Poly(p-xylxlene)<br />

PS Poly(styrene)<br />

PU Poly(urethane)<br />

PVA Poly(vinyl alcohol)<br />

PVAc Poly(vinyl acetate)<br />

P2VP Poly(2-vinyl pyridine)<br />

PVDF Poly(vinylidene fluoride)<br />

PVMS Poly(vinylmethylsiloxane)<br />

PVP Poly(vinylpyrrolidone)<br />

QD Quantum dot<br />

SEM Scanning electron microscopy<br />

TEM Transmission electron microscopy<br />

TUFT Tubes by fiber templates<br />

UV Ultraviolet<br />

38


Figures<br />

Figure 1.1. Schematic figure of basic electrospinning setup. 103<br />

39


Figure 1.2. Photograph of the whipping motion of the instability region of the polymer jet<br />

during the electrospinning process. 12<br />

40


A.<br />

Figure 1.3. A method to produce aligned fibers. A) a schematic utilizing the parallel<br />

grounded electrode collection system and B) the resulting aligned polymer fibers. 25<br />

B.<br />

41


Figure 1.4. SEM micrograph of beaded PEO fibers.<br />

42


Figure 1.5. Reduction in bead density upon increase in polymer solution viscosity. 258<br />

43


Figure 1.6. Ribbon-like fibers formed via electrospinning. 38<br />

44


Figure 1.7. SEM image of anatase hollow fibers created via a co-axial electrospinning<br />

setup. 39<br />

45


Figure 1.8. TEM of PPX/Pd TUFT hybrid nanotubes after the pyrolysis of PLA template<br />

fibers and inset is an electron diffraction pattern of Pd crystals. 46<br />

46


Figure 1.9. SEM micrograph of porous PLA fibers obtained via electrospinning and<br />

subsequent swelling. 259<br />

47


Figure 1.10. SEM iamges of anatase nanofibers whose surfaces have been decorated with a)<br />

gold and b) silver nanoparticles via photocatalytic reduction. 63<br />

48


Figure 1.11. A list of organosoluble polymers and their molecular structure. 103<br />

49


Figure 1.12. Schematic of polymer/inorganic composite nanofibers when a) inorganic ions<br />

are incorporated into electrospun fibers followed by exposure to gas to synthesize inorganic<br />

nanoparticles both inside and outside of the nanofiber and b) when only the surface of<br />

nanofibers are modified with metal ions. 123<br />

52


a<br />

53<br />

100 nm<br />

c<br />

100 nm<br />

Figure 1.13. TEM micrographs of multicomponent polymer electrospun fibers<br />

demonstrating a) A core-sheath structure formed by a polymer blend 128 b) Lamellar structure<br />

formed by a phase-separated block copolymer 260 c) Cylindrical structure formed by a phaseseparated<br />

block copolymer. 260


Figure 1.14. TEM image of a) PAN/CNT composite nanofiber mat and b) demonstrating the<br />

uniform distribution and alignment of CNTs witin a PAN fiber. 56<br />

54


Figure 1.15. Average Young’s modulus for electrosopun nylon-6 and nylon-6/O-MMT<br />

nanocomposite single fibers vs. fiber diameter. 158<br />

55


Figure 1.16. Demonstration that as the aspect ratio of gold nanorods increases, as does the<br />

maximum optical absorbance and thus the color of the aqueous colloidal suspension. 261<br />

56


Figure 1.17. Three-dimensional mineralized electrospun fibers mimicking the hierarchical<br />

structure of bone. 185<br />

57


Figure 1.18. Left: magnetic induction map from two pairs of bacterial magnetite chains.<br />

Right: A bright-field TEM image of a double chain of magnetite magnetosomes. 212<br />

58


Figure 1.19. Pyroloysis of UV cross-linked PS-b-PFEMS films with a) height-mode<br />

scanning force microscopy, b) phase-mode, c) TEM images, and d) A schematic of the<br />

morphology. 262 Inset scale bars = 50 nm.<br />

59


Figure 1.20. TEM micrographs of scarf-shaped PI-b-PFS co-micelles (scale bar = 500<br />

nm). 245<br />

60


References<br />

(1) Rayleigh, L. Philosophy Magazine 1882, 14, 184.<br />

(2) Zeleny, J. Physical Review 1914, 3, 69.<br />

(3) Fenn, J. B.; Mann, M.; Meng, C. K.et al. Science 1989, 246, 64.<br />

(4) Formhals, A. 1934; Vol. US 1,975,504.<br />

(5) Reneker, D. H.; Chun, I. Nanotechnology 1996, 7, 216.<br />

(6) Ondarcuhu, T.; Joachim, C. Europhysics Letters 1998, 42, 215.<br />

(7) J Hagewood, A. W. In Nonwovens World 2003, p 69.<br />

(8) Huang, Z. M.; Zhang, Y. Z.; Kotaki, M.et al. Composites Science and Technology 2003, 63,<br />

2223.<br />

(9) Doshi, J.; Reneker, D. H. Journal of Electrostatics 1995, 35, 151.<br />

(10) Taylor, G. Proceedings of the Royal Society of London Series A 1969, 313, 453.<br />

(11) Shin YM, H. M., Brenner MP Applied Physics Letters 2001, 78, 1149.<br />

(12) Yarin, A. L.; Koombhongse, S.; Reneker, D. H. Journal of Applied Physics 2001, 89, 3018.<br />

(13) Kim JS, R. D. Polymer Engineering and Science 1999, 39, 849.<br />

(14) Theron, S. A.; Yarin, A. L.; Zussman, E.et al. Polymer 2005, 46, 2889.<br />

(15) Li D, X. Y. Nano Letters 2004, 4, 933.<br />

(16) Kalra, V.; Lee, J. H.; Park, J. H.et al. Small 2009, 5, 2323.<br />

(17) Sun, X. Y.; Shankar, R.; Borner, H. G.et al. Advanced Materials 2007, 19, 87.<br />

(18) Kessick, R.; Fenn, J.; Tepper, G. Polymer 2004, 45, 2981.<br />

(19) Yarin, A. L.; Zussman, E. Polymer 2004, 45, 2977.<br />

(20) Sun, D. H.; Chang, C.; Li, S.et al. Nano Letters 2006, 6, 839.<br />

(21) Gonzalez, R.; Pinto, N. J. Synthetic Metals 2005, 151, 275.<br />

61


(22) Huang, Y.; Duan, X. F.; Wei, Q. Q.et al. Science 2001, 291, 630.<br />

(23) Kameoka, J.; Craighead, H. G. Applied Physics Letters 2003, 83, 371.<br />

(24) Theron, A.; Zussman, E.; Yarin, A. L. Nanotechnology 2001, 12, 384.<br />

(25) Li, D.; Xia, Y. N. Nano Letters 2003, 3, 555.<br />

(26) Li, D.; Wang, Y. L.; Xia, Y. N. Advanced Materials 2004, 16, 361.<br />

(27) Lyons, J.; Li, C.; Ko, F. Polymer 2004, 45, 7597.<br />

(28) Thoppey, N. M.; Bochinski, J. R.; Clarke, L. I.et al. Nanotechnology 2011, 22.<br />

(29) Levit, N.; Tepper, G. Journal of Supercritical Fluids 2004, 31, 329.<br />

(30) Um, I. C.; Fang, D. F.; Hsiao, B. S.et al. Biomacromolecules 2004, 5, 1428; Wang, X. F.;<br />

Um, I. C.; Fang, D. F.et al. Polymer 2005, 46, 4853.<br />

(31) Burger, C.; Hsiao, B. S.; Chu, B. In Annual Review of Materials Research 2006; Vol. 36, p<br />

333.<br />

(32) Fong, H.; Reneker, D. H. Journal of Polymer Science Part B-Polymer Physics 1999, 37,<br />

3488.<br />

(33) Zussman, E.; Yarin, A. L.; Weihs, D. Experiments in Fluids 2002, 33, 315.<br />

(34) Liu, H. Q.; Hsieh, Y. L. Journal of Polymer Science Part B-Polymer Physics 2002, 40, 2119.<br />

(35) Deitzel, J. M.; Kleinmeyer, J.; Harris, D.et al. Polymer 2001, 42, 261.<br />

(36) Yao, L.; Haas, T. W.; Guiseppi-Elie, A.et al. Chemistry of Materials 2003, 15, 1860.<br />

(37) Fridrikh, S. V.; Yu, J. H.; Brenner, M. P.et al. Physical Review Letters 2003, 90.<br />

(38) Koombhongse, S.; Liu, W. X.; Reneker, D. H. Journal of Polymer Science Part B-Polymer<br />

Physics 2001, 39, 2598.<br />

(39) Li, D.; Xia, Y. N. Nano Letters 2004, 4, 933.<br />

(40) Li, D.; Xia, Y. N. Advanced Materials 2004, 16, 1151.<br />

62


(41) Shin, Y. M.; Hohman, M. M.; Brenner, M. P.et al. Applied Physics Letters 2001, 78, 1149.<br />

(42) Loscertales, I. G.; Barrero, A.; Marquez, M.et al. Journal of the American Chemical Society<br />

2004, 126, 5376.<br />

(43) Dror, Y.; Kuhn, J.; Avrahami, R.et al. Macromolecules 2008, 41, 4187.<br />

(44) Li, D.; McCann, J. T.; Xia, Y. N. Small 2005, 1, 83.<br />

(45) Dror, Y.; Salalha, W.; Avrahami, R.et al. Small 2007, 3, 1064.<br />

(46) Hou, H. Q.; Jun, Z.; Reuning, A.et al. Macromolecules 2002, 35, 2429.<br />

(47) Caruso, R. A.; Schattka, J. H.; Greiner, A. Advanced Materials 2001, 13, 1577.<br />

(48) Bognitzki, M.; Hou, H. Q.; Ishaque, M.et al. Advanced Materials 2000, 12, 637.<br />

(49) Ochanda, F.; Jones, W. E. Langmuir 2005, 21, 10791.<br />

(50) Reneker, D. H.; Yarin, A. L.; Fong, H.et al. Journal of Applied Physics 2000, 87, 4531.<br />

(51) Samon, J. M.; Schultz, J. M.; Wu, J.et al. Journal of Polymer Science Part B-Polymer Physics<br />

1999, 37, 1277.<br />

(52) Dersch, R.; Liu, T. Q.; Schaper, A. K.et al. Journal of Polymer Science Part a-Polymer<br />

Chemistry 2003, 41, 545.<br />

(53) Kalra, V.; Mendez, S.; Lee, J. H.et al. Advanced Materials 2006, 18, 3299.<br />

(54) Li, D.; Ouyang, G.; McCann, J. T.et al. Nano Letters 2005, 5, 913.<br />

(55) McCann, J. T.; Marquez, M.; Xia, Y. N. Nano Letters 2006, 6, 2868.<br />

(56) Ko, F.; Gogotsi, Y.; Ali, A.et al. Advanced Materials 2003, 15, 1161.<br />

(57) Ra, E. J.; An, K. H.; Kim, K. K.et al. Chem. Phys. Lett. 2005, 413, 188.<br />

(58) Wang, A.; Singh, H.; Hatton, T. A.et al. Polymer 2004, 45, 5505.<br />

(59) Sawicka, K. M.; Gouma, P. Journal of Nanoparticle Research 2006, 8, 769.<br />

(60) Yao, C.; Li, X. S.; Neoh, K. G.et al. Applied Surface Science 2009, 255, 3854.<br />

63


(61) Lin, J.; Qiu, S. Y.; Lewis, K.et al. Biotechnology and Bioengineering 2003, 83, 168.<br />

(62) Deng, J. P.; Wang, L. F.; Liu, L. Y.et al. Progress in Polymer Science 2009, 34, 156.<br />

(63) Li, D.; McCann, J. T.; Gratt, M.et al. Chemical Physics Letters 2004, 394, 387.<br />

(64) Chun, I.; Reneker, D. H.; Fong, H.et al. Journal of Advanced Materials 1999, 31, 36.<br />

(65) Zussman, E.; Chen, X.; Ding, W.et al. Carbon 2005, 43, 2175.<br />

(66) Chen, J. L.; Chu, B.; Hsiao, B. S. Journal of Biomedical Materials Research Part A 2006,<br />

79A, 307.<br />

(67) Muller, K.; Quinn, J. F.; Johnston, A. P. R.et al. Chemistry of Materials 2006, 18, 2397.<br />

(68) Ho, C. C.; Chen, W. S.; Shie, T. Y.et al. Langmuir 2008, 24, 5663.<br />

(69) Luong, N. D.; Moon, I. S.; Lee, D. S.et al. Materials Science & Engineering C-Biomimetic<br />

and Supramolecular Systems 2008, 28, 1242; Dong, F. X.; Li, Z. Y.; Huang, H. M.et al. Materials<br />

Letters 2007, 61, 2556.<br />

(70) Ye, P.; Xu, Z. K.; Wu, J.et al. Biomaterials 2006, 27, 4169.<br />

(71) Wang, Z. G.; Wan, L. S.; Xu, Z. K. Soft Matter 2009, 5, 4161.<br />

(72) Winblade, N. D.; Nikolic, I. D.; Hoffman, A. S.et al. Biomacromolecules 2000, 1, 523.<br />

(73) Kaur, S.; Ma, Z.; Gopal, R.et al. Langmuir 2007, 23, 13085.<br />

(74) Lee, S.; Obendorf, S. K. Textile Research Journal 2007, 77, 696.<br />

(75) Liao, S.; Murugan, R.; Chan, C. K.et al. Journal of the Mechanical Behavior of Biomedical<br />

Materials 2008, 1, 252.<br />

(76) Zeng, J.; Aigner, A.; Czubayko, F.et al. Biomacromolecules 2005, 6, 1484.<br />

(77) Inai, R.; Kotaki, M.; Ramakrishna, S. Nanotechnology 2005, 16, 208.<br />

(78) Lee, K. H.; Kim, H. Y.; La, Y. M.et al. Journal of Polymer Science Part B-Polymer Physics<br />

2002, 40, 2259.<br />

64


(79) Wenger, M. P. E.; Bozec, L.; Horton, M. A.et al. Biophysical Journal 2007, 93, 1255.<br />

(80) Hang, F.; Lu, D.; Bailey, R. J.et al. Nanotechnology 2011, 22.<br />

(81) Riboldi, S. A.; Sampaolesi, M.; Neuenschwander, P.et al. Biomaterials 2005, 26, 4606.<br />

(82) Khil, M. S.; Cha, D. I.; Kim, H. Y.et al. Journal of Biomedical Materials Research Part B-<br />

Applied Biomaterials 2003, 67B, 675.<br />

(83) Ma, Z. W.; Kotaki, M.; Yong, T.et al. Biomaterials 2005, 26, 2527.<br />

(84) Li, J. X.; He, A. H.; Zheng, J. F.et al. Biomacromolecules 2006, 7, 2243.<br />

(85) Morgan, A. W.; Roskov, K. E.; Lin-Gibson, S.et al. Biomaterials 2008, 29, 2556.<br />

(86) Stephens, J. S.; Fahnestock, S. R.; Farmer, R. S.et al. Biomacromolecules 2005, 6, 1405.<br />

(87) Min, B. M.; Jeong, L.; Lee, K. Y.et al. Macromolecular Bioscience 2006, 6, 285.<br />

(88) Jiang, H. L.; Fang, D. F.; Hsiao, B. S.et al. Biomacromolecules 2004, 5, 326.<br />

(89) Min, B. M.; Lee, S. W.; Lim, J. N.et al. Polymer 2004, 45, 7137.<br />

(90) Xie, J. B.; Hsieh, Y. L. Journal of Materials Science 2003, 38, 2125.<br />

(91) Wu, L. L.; Yuan, X. Y.; Sheng, J. Journal of Membrane Science 2005, 250, 167.<br />

(92) Jia, H. F.; Zhu, G. Y.; Vugrinovich, B.et al. Biotechnology Progress 2002, 18, 1027.<br />

(93) McKee, M. G.; Layman, J. M.; Cashion, M. P.et al. Science 2006, 311, 353.<br />

(94) Tan, E. P. S.; Ng, S. Y.; Lim, C. T. Biomaterials 2005, 26, 1453.<br />

(95) Verreck, G.; Chun, I.; Rosenblatt, J.et al. Journal of Controlled Release 2003, 92, 349.<br />

(96) Son, W. K.; Youk, J. H.; Lee, T. S.et al. Polymer 2004, 45, 2959.<br />

(97) Boland, E. D.; Coleman, B. D.; Barnes, C. P.et al. Acta Biomaterialia 2005, 1, 115.<br />

(98) Nair, L. S.; Bhattacharyya, S.; Bender, J. D.et al. Biomacromolecules 2004, 5, 2212.<br />

(99) Liang, D.; Hsiao, B. S.; Chu, B. Advanced Drug Delivery Reviews 2007, 59, 1392.<br />

65


(100) Katti, D. S.; Robinson, K. W.; Ko, F. K.et al. Journal of Biomedical Materials Research Part<br />

B-Applied Biomaterials 2004, 70B, 286.<br />

(101) Tomczak, N.; van Hulst, N. F.; Vancso, G. J. Macromolecules 2005, 38, 7863.<br />

(102) Koski, A.; Yim, K.; Shivkumar, S. Materials Letters 2004, 58, 493.<br />

(103) Greiner, A.; Wendorff, J. H. Angewandte Chemie-International Edition 2007, 46, 5670.<br />

(104) Kumeta, K.; Nagashima, I.; Matsui, S.et al. Journal of Applied Polymer Science 2003, 90,<br />

2420.<br />

(105) Li, L.; Hsieh, Y. L. Polymer 2005, 46, 5133.<br />

(106) Zeng, J.; Hou, H. Q.; Wendorff, J. H.et al. Macromolecular Rapid Communications 2005, 26,<br />

1557.<br />

(107) Bergshoef, M. M.; Vancso, G. J. Advanced Materials 1999, 11, 1362.<br />

(108) Sutasinpromprae, J.; Jitjaicham, S.; Nithitanakul, M.et al. Polymer International 2006, 55,<br />

825.<br />

(109) Zhang, D.; Karki, A. B.; Rutman, D.et al. Polymer 2009, 50, 4189.<br />

(110) Ge, J. J.; Hou, H. Q.; Li, Q.et al. Journal of the American Chemical Society 2004, 126,<br />

15754.<br />

(111) Katta, P.; Alessandro, M.; Ramsier, R. D.et al. Nano Letters 2004, 4, 2215.<br />

(112) Suthar, A.; Chase, G. Tce 2001, 26.<br />

(113) Chronakis, I. S.; Jakob, A.; Hagstrom, B.et al. Langmuir 2006, 22, 8960.<br />

(114) Kim, J. R.; Choi, S. W.; Jo, S. M.et al. Electrochimica Acta 2004, 50, 69.<br />

(115) Cha, D. I.; Kim, H. Y.; Lee, K. H.et al. Journal of Applied Polymer Science 2005, 96, 460.<br />

(116) Hench, L. L.; West, J. K. Chemical Reviews 1990, 90, 33.<br />

66


(117) Larsen, G.; Velarde-Ortiz, R.; Minchow, K.et al. Journal of the American Chemical Society<br />

2003, 125, 1154.<br />

(118) Wang, Y.; Furlan, R.; Ramos, I.et al. Applied Physics a-Materials Science & Processing<br />

2004, 78, 1043.<br />

(119) Choi, S. S.; Lee, S. G.; Im, S. S.et al. Journal of Materials Science Letters 2003, 22, 891.<br />

(120) Madhugiri, S.; Sun, B.; Smirniotis, P. G.et al. Microporous and Mesoporous Materials 2004,<br />

69, 77.<br />

(121) Bai, J.; Li, Y. X.; Yang, S. T.et al. Nanotechnology 2007, 18.<br />

(122) Lu, X. F.; Zhao, Y. Y.; Wang, C. Advanced Materials 2005, 17, 2485.<br />

(123) Lu, X. F.; Zhao, Y. J.; Wang, C.et al. Macromolecular Rapid Communications 2005, 26,<br />

1325.<br />

(124) Wang, W.; Asher, S. A. Journal of the American Chemical Society 2001, 123, 12528.<br />

(125) Son, W. K.; Youk, J. H.; Lee, T. S.et al. Macromolecular Rapid Communications 2004, 25,<br />

1632.<br />

(126) Jin, H. J.; Fridrikh, S.; Rutledge, G. C.et al. Abstracts of Papers of the American Chemical<br />

Society 2002, 224, U431.<br />

(127) Mitchell, S. B.; Sanders, J. E. Journal of Biomedical Materials Research Part A 2006, 78A,<br />

110.<br />

(128) Wei, M.; Lee, J.; Kang, B. W.et al. Macromolecular Rapid Communications 2005, 26, 1127.<br />

(129) Kameoka, J.; Ilic, R.; Czaplewski, D.et al. Journal of Photopolymer Science and Technology<br />

2004, 17, 421.<br />

(130) Jiang, H. L.; Hu, Y. Q.; Li, Y.et al. Journal of Controlled Release 2005, 108, 237.<br />

67


(131) Greiner, A.; Wendorff, J. H.; Yarin, A. L.et al. Applied Microbiology and Biotechnology<br />

2006, 71, 387.<br />

(132) Valiquette, D.; Pellerin, C. Macromolecules 2011, 44, 2838.<br />

(133) Chen, M. L.; Dong, M. D.; Havelund, R.et al. Chemistry of Materials 2010, 22, 4214.<br />

(134) Zhang, Y. Z.; Huang, Z. M.; Xu, X. J.et al. Chemistry of Materials 2004, 16, 3406.<br />

(135) Ristolainen, N.; Heikkila, P.; Harlin, A.et al. Macromolecular Materials and Engineering<br />

2006, 291, 114.<br />

(136) Kim, J. S.; Lee, D. S. Polymer Journal 2000, 32, 616.<br />

(137) Greiner, A.; Wendorff, J. H. Angewandte Chemie 2007, 46, 5670.<br />

(138) Ma, M.; Krikorian, V.; Yu, J. H.et al. Nano Letters 2006, 6, 2969; Kalra, V.; Kakad, P. A.;<br />

Mendez, S.et al. Macromolecules 2006, 39, 5453.<br />

(139) Ma, M. L.; Hill, R. M.; Lowery, J. L.et al. Langmuir 2005, 21, 5549; Alli, A.; Hazer, B.;<br />

Menceloglu, Y.et al. European Polymer Journal 2006, 42, 740.<br />

(140) Kim, T. G.; Park, T. G. Biotechnology Progress 2006, 22, 1108.<br />

(141) Luu, Y. K.; Kim, K.; Hsiao, B. S.et al. Journal of Controlled Release 2003, 89, 341.<br />

(142) Baughman, R. H.; Zakhidov, A. A.; de Heer, W. A. Science 2002, 297, 787.<br />

(143) Salvetat, J. P.; Bonard, J. M.; Thomson, N. H.et al. Applied Physics a-Materials Science &<br />

Processing 1999, 69, 255.<br />

(144) Zhang, Y.; Lu, F.; Wang, Z.et al. Crystal Growth & Design 2007, 7, 1459; Ayutsede, J.;<br />

Gandhi, M.; Sukigara, S.et al. Biomacromolecules 2006, 7, 208.<br />

(145) Kannan, P.; Eichhorn, S. J.; Young, R. J. Nanotechnology 2007, 18.<br />

(146) Kim, G. M.; Michler, G. H.; Potschke, P. Polymer 2005, 46, 7346.<br />

(147) Jeong, J. S.; Jeon, S. Y.; Lee, T. Y.et al. Diamond and Related Materials 2006, 15, 1839.<br />

68


(148) Sen, R.; Zhao, B.; Perea, D.et al. Nano Letters 2004, 4, 459.<br />

(149) Dror, Y.; Salalha, W.; Khalfin, R. L.et al. Langmuir 2003, 19, 7012.<br />

(150) Sung, J. H.; Kim, H. S.; Jin, H. J.et al. Macromolecules 2004, 37, 9899.<br />

(151) Li, J. X.; He, A. H.; Han, C. C.et al. Macromolecular Rapid Communications 2006, 27, 114.<br />

(152) Zhong, S. P.; Teo, W. E.; Zhu, X.et al. Biomacromolecules 2005, 6, 2998.<br />

(153) Kwon, I. K.; Matsuda, T. Biomacromolecules 2005, 6, 2096.<br />

(154) Li, C. M.; Vepari, C.; Jin, H. J.et al. Biomaterials 2006, 27, 3115.<br />

(155) Sawicka, K. M.; Prasad, A. K.; Gouma, P. I. Sensor Letters 2005, 3, 31.<br />

(156) Ren, G.; Xu, X.; Liu, Q.et al. Reactive & Functional Polymers 2006, 66, 1559.<br />

(157) Teo, W.-E.; Ramakrishna, S. Composites Science and Technology 2009, 69, 1804.<br />

(158) Li, L.; Bellan, L. M.; Craighead, H. G.et al. Polymer 2006, 47, 6208.<br />

(159) Hwang, J.; Muth, J.; Ghosh, T. Journal of Applied Polymer Science 2007, 104, 2410.<br />

(160) Tiwari, M. K.; Yarin, A. L.; Megaridis, C. M. Journal of Applied Physics 2008, 103.<br />

(161) Wutticharoenmongkol, P.; Sanchavanakit, N.; Pavasant, P.et al. Macromolecular Bioscience<br />

2006, 6, 70.<br />

(162) Bao, Q.; Zhang, H.; Yang, J.-x.et al. Advanced Functional Materials 2010, 20, 782.<br />

(163) Sui, X.; Shao, C.; Liu, Y. Polymer 2007, 48, 1459.<br />

(164) Li, M.; Zhang, J.; Zhang, H.et al. Advanced Functional Materials 2007, 17, 3650.<br />

(165) Peresin, M. S.; Habibi, Y.; Zoppe, J. O.et al. Biomacromolecules 2010, 11, 674.<br />

(166) Wang, S.; Li, Y.; Fei, X.et al. Journal of Colloid and Interface Science 2011, 359, 380.<br />

(167) Maity, S.; Downen, L. N.; Bochinski, J. R.et al. Polymer 2011, 52, 1674.<br />

(168) Kim, G.-M.; Wutzler, A.; Radusch, H.-J.et al. Chemistry of Materials 2005, 17, 4949.<br />

69


(169) Jin, W. J.; Lee, H. K.; Jeong, E. H.et al. Macromolecular Rapid Communications 2005, 26,<br />

1903.<br />

(170) Yu, N.; Shao, C.; Liu, Y.et al. Journal of Colloid and Interface Science 2004, 285, 163.<br />

(171) Hong, K. H.; Park, J. L.; Sul, I. H.et al. Journal of Polymer Science Part B-Polymer Physics<br />

2006, 44, 2468.<br />

(172) Kelly, K. L.; Coronado, E.; Zhao, L. L.et al. J. Phys. Chem. B 2003, 107, 668.<br />

(173) Link, S.; El-Sayed, M. A. J. Phys. Chem. B 1999, 103, 8410.<br />

(174) Demir, M. M.; Gulgun, M. A.; Menceloglu, Y. Z.et al. Macromolecules 2004, 37, 1787.<br />

(175) Wang, M.; Singh, H.; Hatton, T. A.et al. Polymer 2004, 45, 5504.<br />

(176) Zhu, J. H.; Wei, S. Y.; Chen, X. L.et al. Journal of Physical Chemistry C 2010, 114, 8844.<br />

(177) Pariente, J. L.; Kim, B. S.; Atala, A. Journal of Biomedical Materials Research 2001, 55, 33.<br />

(178) Xu, C. Y.; Yang, F.; Wang, S.et al. Journal of Biomedical Materials Research Part A 2004,<br />

71A, 154.<br />

(179) Min, B. M.; Lee, G.; Kim, S. H.et al. Biomaterials 2004, 25, 1289; Lee, I. S.; Kwon, O. H.;<br />

Meng, W.et al. Macromolecular Research 2004, 12, 374.<br />

(180) Boudriot, U.; Dersch, R.; Goetz, B.et al. Biomedizinische Technik 2004, 49, 242.<br />

(181) Zhang, Y. Z.; Venugopal, J.; Huang, Z. M.et al. Biomacromolecules 2005, 6, 2583; Ghasemi-<br />

Mobarakeh, L.; Prabhakaran, M. P.; Morshed, M.et al. Biomaterials 2008, 29, 4532.<br />

(182) Fujihara, K.; Kotaki, M.; Ramakrishna, S. Biomaterials 2005, 26, 4139.<br />

(183) Thomas, V.; Zhang, X.; Catledge, S. A.et al. Biomedical Materials 2007, 2, 224.<br />

(184) Schnell, E.; Klinkhammer, K.; Balzer, S.et al. Biomaterials 2007, 28, 3012.<br />

(185) Teo, W. E.; Liao, S.; Chan, C. K.et al. Current Nanoscience 2008, 4, 361.<br />

70


(186) Park, S. J.; Kang, Y. C.; Park, J. Y.et al. Journal of <strong>Engineered</strong> Fibers and Fabrics 2010, 5,<br />

50.<br />

(187) Kobayashi, S.; Akiyama, R. Chemical Communications 2003, 449; Stasiak, M.; Studer, A.;<br />

Greiner, A.et al. Chemistry-a European Journal 2007, 13, 6150.<br />

(188) Li, M. Y.; Han, G. Y.; Yang, B. S. Electrochemistry Communications 2008, 10, 880.<br />

(189) Nie, H.; Soh, B. W.; Fu, Y. C.et al. Biotechnology and Bioengineering 2008, 99, 223;<br />

Sikareepaisan, P.; Suksamrarn, A.; Supaphol, P. Nanotechnology 2008, 19.<br />

(190) Muller, R. H.; Mader, K.; Gohla, S. European Journal of Pharmaceutics and<br />

Biopharmaceutics 2000, 50, 161; Soppimath, K. S.; Aminabhavi, T. M.; Kulkarni, A. R.et al. Journal<br />

of Controlled Release 2001, 70, 1.<br />

(191) Xie, J. W.; Wang, C. H. Pharmaceutical Research 2006, 23, 1817.<br />

(192) Sanders, E. H.; Kloefkorn, R.; Bowlin, G. L.et al. Macromolecules 2003, 36, 3803.<br />

(193) Aussawasathien, D.; Teerawattananon, C.; Vongachariya, A. Journal of Membrane Science<br />

2008, 315, 11.<br />

(194) Desai, K.; Kit, K.; Li, J.et al. Biomacromolecules 2008, 9, 1000.<br />

(195) Ki, C. S.; Gang, E. H.; Um, N. C.et al. Journal of Membrane Science 2007, 302, 20.<br />

(196) Son, W. K.; Youk, J. H.; Park, W. H. Carbohydrate Polymers 2006, 65, 430; Yao, C.; Li, X.<br />

S.; Neoh, K. G.et al. Journal of Membrane Science 2008, 320, 259.<br />

(197) Kenawy, E. R.; Mahmoud, Y. A. G. Macromolecular Bioscience 2003, 3, 107.<br />

(198) Tan, K.; Obendorf, S. K. Journal of Membrane Science 2007, 305, 287.<br />

(199) Hota, G.; Kumar, B. R.; Ramakrishna, W. Journal of Materials Science 2008, 43, 212.<br />

(200) Zhang, L. F.; Luo, J. E.; Menkhaus, T. J.et al. Journal of Membrane Science 2011, 369, 499.<br />

(201) Kaur, S.; Gopal, R.; Ng, W. J.et al. Mrs Bulletin 2008, 33, 21.<br />

71


(202) Ramaseshan, R.; Sundarrajan, S.; Jose, R.et al. Journal of Applied Physics 2007, 102.<br />

(203) Ramaseshan, R.; Sundarrajan, S.; Liu, Y. J.et al. Nanotechnology 2006, 17, 2947.<br />

(204) Kraybill, H. F. Bulletin of the New York Academy of Medicine 1978, 54, 413.<br />

(205) Rakitin, A.; Aich, P.; Papadopoulos, C.et al. Physical Review Letters 2001, 86, 3670.<br />

(206) Tanaka, K.; Tengeiji, A.; Kato, T.et al. Science 2003, 299, 1212.<br />

(207) Jiles, D. Introduction to magnetism and magnetic materials; Chapman and Hall: Boca Raton,<br />

1998; Vol. 2.<br />

(208) Yuan, J.; Xu, Y.; Mueller, A. H. E. Chemical Society Reviews 2011, 40, 640.<br />

(209) Lorcy, J. M.; Massuyeau, F.; Moreau, P.et al. Nanotechnology 2009, 20.<br />

(210) Yan, X. H.; Liu, G. J.; Haeussler, M.et al. Chemistry of Materials 2005, 17, 6053; Yan, X. H.;<br />

Liu, G. J.; Liu, F. T.et al. Angewandte Chemie-International Edition 2001, 40, 3593.<br />

(211) Zhang, M. F.; Estournes, C.; Bietsch, W.et al. Advanced Functional Materials 2004, 14, 871.<br />

(212) Simpson, E. T.; Kasama, T.; Posfai, M.et al. In Fifth International Conference on Fine<br />

Particle Magnetism; Pankhurst, Q., Ed. 2005; Vol. 17, p 108.<br />

(213) Graeser, M.; Bognitzki, M.; Massa, W.et al. Advanced Materials 2007, 19, 4244.<br />

(214) Bangar, M. A.; Hangarter, C. M.; Yoo, B.et al. Electroanalysis 2009, 21, 61.<br />

(215) Corr, S. A.; Byrne, S. J.; Tekoriute, R.et al. Journal of the American Chemical Society 2008,<br />

130, 4214.<br />

(216) Park, J. H.; von Maltzahn, G.; Zhang, L. L.et al. Advanced Materials 2008, 20, 1630.<br />

(217) Nguyen P, G.-E. P., Manners I Chemistry Reviews 1999, 99, 1515.<br />

(218) Manners, I. Science 2001, 294, 1664.<br />

(219) Arimoto FS, H. A. Journal of the American Chemical Society 1955, 77, 6295.<br />

72


(220) Abd-el-Aziz, A. S.; Manners, I. Frontiers in Transition Metal-Containing Polymers; Wiley<br />

VCH: Hoboken, 2007.<br />

(221) Han, F.-S.; Higuchi, M.; Kurth, D. G. Tetrahedron 2008, 64, 9108.<br />

(222) Lidrissi, C.; Romerosa, A.; Saoud, M.et al. Angewandte Chemie-International Edition 2005,<br />

44, 2568.<br />

(223) Bender, J. L.; Corbin, P. S.; Fraser, C. L.et al. Journal of the American Chemical Society<br />

2002, 124, 8526; Fustin, C.-A.; Guillet, P.; Schubert, U. S.et al. Advanced Materials 2007, 19, 1665.<br />

(224) Fustin, C. A.; Lohmeijer, B. G. G.; Duwez, A. S.et al. Advanced Materials 2005, 17, 1162.<br />

(225) Leong, W. L.; Tam, A. Y.-Y.; Batabyal, S. K.et al. Chemical Communications 2008, 3628.<br />

(226) Xing, B. G.; Choi, M. F.; Xu, B. Chemistry-a European Journal 2002, 8, 5028; Miravet, J. F.;<br />

Escuder, B. Chemical Communications 2005, 5796.<br />

(227) Wei, Q.; James, S. L. Chemical Communications 2005, 1555; Yin, J.; Yang, G.; Wang, H.et<br />

al. Chemical Communications 2007, 4614.<br />

(228) Arsenault, A. C.; Miguez, H.; Kitaev, V.et al. Advanced Materials 2003, 15, 503.<br />

(229) Foucher DA, T. B.-Z., Manners I Journal of the American Chemical Society 1992, 114, 6246.<br />

(230) Manners, I. Candian Journal of Chemistry 1998, 76, 371; Resendes R, N. J., Fischer A, et al.<br />

Journal of the American Chemical Society 2001, 123, 2116.<br />

(231) Rulkens, R.; Lough, A. J.; Lovelace, S. R.et al. Journal of the American Chemical Society<br />

1996, 118, 12683.<br />

(232) Manners, I. Synthetic Metal-Containing Polymers; Wiley-VCH: Weinheim, 2004.<br />

(233) Kulbaba, K.; Manners, I. Macromolecular Rapid Communications 2001, 22, 711.<br />

(234) McDowell, J. J.; Zacharia, N. S.; Puzzo, D.et al. Journal of the American Chemical Society<br />

2010, 132, 3236.<br />

73


(235) IW, H. The Physics of Block Copolymers Oxford, UK, 1998.<br />

(236) Hadjichristidis, N.; Pitsikalis, M.; Iatrou, H.et al. Macromolecular Rapid Communications<br />

2003, 24, 979.<br />

(237) Ni, Y. Z.; Rulkens, R.; Manners, I. Journal of the American Chemical Society 1996, 118,<br />

4102.<br />

(238) Schubert, U. S.; Eschbaumer, C. Angewandte Chemie-International Edition 2002, 41, 2893.<br />

(239) Cao, L.; Manners, I.; Winnik, M. A. Macromolecules 2002, 35, 8258.<br />

(240) Wang, X. S.; Winnik, M. A.; Manners, I. Angewandte Chemie-International Edition 2004,<br />

43, 3703.<br />

(241) Rider, D. A.; Cavicchi, K. A.; Power-Billard, K. N.et al. Macromolecules 2005, 38, 6931.<br />

(242) Kim, K. T.; Vandermeulen, G. W. M.; Winnik, M. A.et al. Macromolecules 2005, 38, 4958.<br />

(243) Kloninger, C.; Rehahn, M. Macromolecules 2004, 37, 8319; Lammertink, R.; Hempenius,<br />

M.; Thomas, E.et al. Journal of Polymer Science: Part B: Polymer Physics 1998, 37, 1009; Temple,<br />

K.; Kulbaba, K.; Power-Billard, K. N. Advanced Materials 2003, 15, 297.<br />

(244) Yusoff, S. F. M.; Gilroy, J. B.; Cambridge, G.et al. Journal of the American Chemical Society<br />

2011, 133, 11220.<br />

(245) Gaedt, T.; Ieong, N. S.; Cambridge, G.et al. Nature Materials 2009, 8, 144.<br />

(246) Wang, H.; Patil, A. J.; Liu, K.et al. Advanced Materials 2009, 21, 1805; Wang, H.; Wang, X.;<br />

Winnik, M. A.et al. Journal of the American Chemical Society 2008, 130, 12921.<br />

(247) He, F.; Gadt, T.; Manners, I.et al. Journal of the American Chemical Society 2011, 133, 9095.<br />

(248) Lastella, S.; Jung, Y. J.; Yang, H. C.et al. Journal of Materials Chemistry 2004, 14, 1791; Lu,<br />

J. Q.; Kopley, T. E.; Moll, N.et al. Chemistry of Materials 2005, 17, 2227.<br />

(249) Temple, K.; Kulbaba, K.; Power-Billard, K. N.et al. Advanced Materials 2003, 15, 297.<br />

74


(250) Hinderling, C.; Keles, Y.; Stockli, T.et al. Advanced Materials 2004, 16, 876.<br />

(251) Massey, J. A.; Power, K. N.; Winnik, M. A.et al. Advanced Materials 1998, 10, 1559.<br />

(252) Raez, J.; Manners, I.; Winnik, M. A. Journal of the American Chemical Society 2002, 124,<br />

10381.<br />

(253) Raez, J.; Manners, I.; Winnik, M. A. Langmuir 2002, 18, 7229.<br />

(254) Massey JA, T. K., Cao L, et al. Journal of the American Chemical Society 2000, 122, 11577.<br />

(255) Wang, H.; Winnik, M. A.; Manners, I. Macromolecules 2007, 40, 3784.<br />

(256) Wang, X.; Liu, K.; Aresenault, A. C.et al. Journal of the American Chemical Society 2007,<br />

129, 5630.<br />

(257) Wang, X. S.; Wang, H.; Coombs, N.et al. Journal of the American Chemical Society 2005,<br />

127, 8924.<br />

(258) Fong, H.; Chun, I.; Reneker, D. H. Polymer 1999, 40, 4585.<br />

(259) Bognitzki, M.; Czado, W.; Frese, T.et al. Advanced Materials 2001, 13, 70.<br />

(260) Ma, M. L.; Krikorian, V.; Yu, J. H.et al. Nano Letters 2006, 6, 2969.<br />

(261) Mulvaney, P.; Perez-Juste, J.; Giersig, M.et al. Plasmonics 2006, 1, 61.<br />

75


Abstract<br />

CHAPTER II<br />

Long-Range Alignment of Gold Nanorods in Electrospun Polymer<br />

Nano/Microfibers*<br />

In this study, a scalable fabrication technique for controlling and maintaining the<br />

nanoscale orientation of gold nanorods (GNRs) with long-range macroscale order has been<br />

achieved through electrospinning. The volume fraction of GNRs with an average aspect ratio<br />

of 3.1 is varied from 0.006 to 0.045 in aqueous poly(ethylene oxide) solutions to generate<br />

electrospun fibers possessing different GNR concentrations and measuring 40-3000 nm in<br />

diameter. The GNRs within these fibers exhibit excellent alignment with their longitudinal<br />

axis parallel to the fiber axis n. According to microscopy analysis, the average deviant angle<br />

between the GNR axis and n increases modestly from 3.8 to 13.3° as the fiber diameter<br />

increases. Complementary electron diffraction measurements confirm preferred orientation<br />

of the {100} and {111} GNR planes. Optical absorbance spectroscopy measurements reveal<br />

that the longitudinal surface plasmon resonance bands of the aligned GNRs depend on the<br />

polarization angle and that maximum absorption occurs when the polarization is parallel to n.<br />

*This chapter has been published in its entirety:<br />

KE Roskov, KA Kozek, WC Wu, RK Chhetri, AL Oldenburg, RJ Spontak, JB Tracy. “Long-<br />

Range Alignment of Gold Nanorods in Electrospun Polymer Nano/Microfibers.” Langmuir. Web<br />

publication date: 8-11-11, DOI # 10.1021/la202106<br />

76


Nanoparticle (NP) synthesis, characterization and self-assembly have been thoroughly<br />

investigated over the past two decades primarily because of their novel size- and shape-<br />

tunable functional properties, 1 as well as the wide variety of property enhancements they<br />

impart to matrix materials, such as polymers. 2 As their size decreases, metal NPs can exhibit<br />

physical and chemical properties (e.g., catalytic, 3 optical, 4 electronic, 5 and magnetic 6 ) that are<br />

not observed in the bulk. Gold and silver NPs, for instance, are known to exhibit strong<br />

surface plasmon resonances in the visible electromagnetic spectrum as a consequence of<br />

optically-driven coherent oscillations of conduction electrons. 7 These resonances give rise to<br />

characteristic optical absorption and scattering spectra, 8 yielding brightly colored, NP-<br />

containing suspensions. Gold nanorods 9 (GNRs) have recently attracted considerable<br />

attention due to their strong and uniquely tunable longitudinal surface plasmon resonance<br />

(LSPR) along the longitudinal GNR axis. The LSPR wavelength of GNRs can be adjusted<br />

from ~520 nm in the case of low-aspect-ratio (i.e., spherical) gold NPs to at least 1750 nm in<br />

the near-infrared region of the electromagnetic spectrum for high-aspect-ratio GNRs. 10<br />

Currently, GNRs with a LSPR at 800 nm are of particularly significant interest for in vivo<br />

imaging of biological systems, because blood and tissue exhibit an absorbance minimum at<br />

this wavelength. 11<br />

When dispersed in a solvent, GNRs are randomly oriented, and they tend to remain so<br />

when physically deposited on surfaces. If the distribution of GNR sizes and shapes is<br />

sufficiently narrow so that the GNRs can be considered nearly monodisperse, they can<br />

spontaneously self-assemble into ordered grains with different orientations. 12 Presently,<br />

77


alignment of such grains over macroscale dimensions is a significant technological challenge,<br />

but is highly desirable, 13 because the optical anisotropy 14 of GNRs at the LSPR can be<br />

greater than ~250:1. 15 A scalable method for controlling and maintaining the orientation of<br />

GNRs at the nanoscale, while fabricating macroscale structures that exploit GNR orientation,<br />

is therefore needed for creating nanostructured composites with tunable, anisotropic optical<br />

properties. Alignment-enhanced absorption of polymer/GNR composites has been previously<br />

reported using different preparation strategies: thin-film stretching, 16 block copolymer-<br />

templated organization, 17 directed nanoscale assembly, 18 ultrathin film confinement, 19 and<br />

polymer fiber coating. 20 An ongoing challenge for polymer/GNR composites produced by<br />

these methods is the consistent acquisition of long-range, scalable order of highly-aligned<br />

GNRs along a common axis (rather than randomly oriented in a common plane). Here, we<br />

describe a general technique for aligning GNRs within electrospun polymer nano/microfibers<br />

with diameters ranging from 40 to 3000 nm. Our approach enables the hierarchical alignment<br />

of fibers containing aligned nanorods over large, macroscopic dimensions. The resultant<br />

polymer/GNR composites were characterized by electron microscopy, electron diffraction,<br />

and optical absorbance spectroscopy.<br />

The GNRs investigated here were synthesized by modifying a method originally<br />

introduced by Nikoobakht et al. 21 Details of the GNR synthesis are provided as Supporting<br />

Information. A custom-built electrospinning unit with an Al collection target was operated at<br />

an electric potential of 10 kV with a plate distance of 15 cm and solution flow rates of 30-50<br />

�L/min. Aligned electrospun fibers, rather than conventional, randomly oriented fiber mats,<br />

78


were generated by electrospinning between two grounded electrodes separated by 2 cm. 22<br />

Prior to electrospinning, the GNR stock suspension was warmed with a heat gun to<br />

redissolve precipitated cetyltrimethylammonium bromide (CTAB). Poly(ethylene oxide)<br />

(PEO: 1000 kDa, Sigma-Aldrich) was selected for its intrinsic hydrophilicity and, by<br />

inference, its compatibility with the native CTAB coating on the GNRs. A high-molecular-<br />

weight grade was chosen to facilitate fiber formation and impart mechanical robustness (for<br />

handling purposes). While other matrix materials have not yet been explored, we anticipate<br />

similar results would be obtained for comparably water-soluble polymers amenable to<br />

electrospinning. The PEO was dissolved directly into the GNR stock suspension to yield a<br />

maximum GNR volume fraction (�) of 0.045. Lower GNR volume fractions were<br />

subsequently prepared by first dissolving the polymer in a separate deionized water solution<br />

and then adding it to the GNR suspension. Polymer concentrations ranged between 3.0 and<br />

4.5 wt % PEO. Immediately before electrospinning, the mixture was sonicated for 10 min to<br />

ensure a uniform dispersion. Electrospun fibers were dried under vacuum for 24 h at ambient<br />

temperature.<br />

Specimens imaged by transmission electron microscopy (TEM) consisted of randomly<br />

oriented fibers electrospun directly onto carbon-coated Cu grids. Images and selected-area<br />

electron diffraction (SAED) patterns were acquired on a field-emission Hitachi HF2000<br />

microscope operated at 200 kV. Corresponding scanning electron microscopy (SEM) images<br />

were collected from specimens sputter-coated with 6 nm of Au/Pd on a JEOL 6400F field-<br />

emission microscope operated at 5 kV. For optical characterization, a custom-built polarized<br />

79


UV-vis spectrophotometer, housing a fiber-coupled white light source (Photon Control),<br />

broadband linear polarizer, light collection fiber coupled into a USB spectrometer<br />

(Lightspeed Technologies), and a series of collimation and collection lenses, was employed.<br />

Absorbance spectra were acquired by orienting each specimen normal to the collimated white<br />

light beam, which was passed through a ~2 mm diameter aperture and the polarizer.<br />

The GNRs employed in this study have an average width and length of 17 ± 6 and 49 ± 10<br />

nm, respectively (Figure 1), thereby yielding an average aspect (length/diameter) ratio of 3.1.<br />

Electrospinning is becoming established as an important method for preparing polymer<br />

nano/microfibers, because it is straightforward to perform and offers the flexibility to tune<br />

fiber characteristics by controllably varying solution and/or processing parameters. Aligned<br />

fibers generated by electrospinning between two grounded electrodes yield a freestanding,<br />

oriented mat measuring 10 cm wide and 3 cm or more long (Figure 2). Here and in<br />

subsequent discussion, n refers to the fiber axis direction. The fiber thickness is governed by<br />

the PEO concentration when all other electrospinning parameters are held constant. For<br />

instance, the nanofibers shown in Figure 3a with an average diameter of ~50 nm were<br />

electrospun from an aqueous suspension with � = 0.006 and a PEO concentration of 3.2 wt<br />

%. At this relatively low GNR loading in PEO nanofibers, the GNRs orient with their<br />

longitudinal axis parallel to n. To quantify the extent of alignment along n, the average<br />

deviant angle (), determined from the angular difference between the GNR orientation<br />

and n, has been measured for 150-300 GNRs embedded within electrospun PEO fibers of<br />

varying thickness. For nanofibers such as those shown in Figure 3a, = 3.8°. Due to the<br />

80


nanoscale diameter of the PEO fibers and their CTAB surface coating, 9 very few GNR<br />

aggregates are observed at this low loading level, for which the interparticle distance is<br />

typically much longer than the length of the GNR. The theory proposed by Bates and<br />

Frenkel 23 for hard-rod fluids predicts that GNRs with aspect ratios < 7 should exhibit random<br />

orientation in solution, insofar as the GNRs do not interact with each other over large<br />

distances. According to our measurements, it follows that oriented, well-separated GNRs are<br />

aligned by forces other than those arising from interparticle interactions.<br />

During electrospinning, the viscous polymer/GNR solution at the tip of the syringe can be<br />

modeled 24 as a fluid cone leading to a jet, where the charged polymer solution is emitted.<br />

Sink-like flow emerges at the apex of the cone, and streamlines form due to the rapid<br />

decrease in area. In suspensions, the GNRs are initially randomly oriented but begin to align<br />

along the streamlines leading to the jet. Since the Reynolds number, defined as the ratio of<br />

inertial (drag) to viscous forces, is much less than unity at this point, the translational<br />

velocity component dominates, 25 in which case the center of each GNR experiences the same<br />

velocity as the local fluid velocity. As the polymer solution leaves the jet, the GNRs are<br />

expected to be oriented for the most part along the fiber spinning direction. Indeed, the<br />

alignment of significantly longer carbon nanotubes 26 and CdS nanorods with an aspect ratio<br />

of ~20 27 have been experimentally confirmed in electrospun fibers, but we are unaware of<br />

prior studies of nanorods with shorter aspect ratios. When the GNR loading is increased to Φ<br />

= 0.045 (Figure 3b), the GNRs remain highly oriented, because their orientation is<br />

predominately dictated by the local velocity profile. For an increased fiber diameter of ~650<br />

81


nm and Φ = 0.035 (Figure 3c), the velocity profile is forced to extend across the diameter of<br />

the microfiber. Consequently, the GNR longitudinal axes remain highly aligned with n, but<br />

increases modestly to 8.8°. When the fiber diameter is further increased to ~3000 nm<br />

and Φ = 0.031 (Figure 3d), the degree of GNR alignment decreases, as verified by an<br />

accompanying increase in to 13.3°.<br />

Two other key morphological observations warrant mention. The first is that long GNRs<br />

are more highly oriented with respect to n than are short ones. For example, in a fiber<br />

measuring ~80 nm in diameter, when measurements of are sorted according to the<br />

GNR length, the shortest 26% of the GNRs possess that is 54% greater than for<br />

all GNRs in the fiber. Secondly, the fiber diameter is one of the main factors that determine<br />

the degree of GNR alignment within electrospun fibers. That is, within fibers of nearly<br />

constant diameter, the extent of alignment does not appear to depend strongly on Φ over the<br />

range of GNR concentrations investigated. These two results imply that the degree of<br />

nanoscale GNR orientation over macroscale dimensions is primarily controlled by the flow<br />

field introduced by the polymer and experienced by the GNRs during electrospinning,<br />

coupled with the GNR aspect ratio.<br />

Selected-area electron diffraction (SAED) performed on a segment of microfiber<br />

measuring 200 nm in diameter and containing ~20 GNRs yields the pattern included as an<br />

inset in Figure 3b. Previous studies have shown that GNRs possess a faceted crystal structure<br />

with a [110] growth direction, {111} end facets, and {100} side facets. 28 Analysis of the<br />

SAED pattern in Figure 3b confirms the existence of a face-centered cubic lattice with<br />

82


preferred orientations identified by the {100} and {111} reflections, which is consistent with<br />

results reported elsewhere. 21 These reflections appear over a limited angular range rather than<br />

as single spots, which indicates a distribution of GNR orientations. This result has been<br />

analyzed in terms of the truncated Herman’s orientation function (P2), 29 written as<br />

��<br />

P2 � 3 cos2 � �1<br />

. (1)<br />

2<br />

Here, � is the azimuthal angle extending from 0 to 2� around the circular SAED pattern, and<br />

cos 2 � =<br />

I(�)cos 2 � (�)d�<br />

, (2)<br />

� I(�)d�<br />

where I(�) is the scattering intensity that varies along a circular trace of constant radius.<br />

��<br />

Limiting values of<br />

��<br />

P2 are 1.0 for perfect alignment along n, 0.0 for random dispersions and<br />

�0.5 for perpendicular alignment relative to n. The present analysis yields P2 = 0.73 for the<br />

aligned GNR-containing PEO microfibers under investigation, which further confirms that<br />

the GNRs are highly oriented within the microfibers.<br />

Gold nanorods are of particular contemporary interest because of their anisotropic optical<br />

properties. For observing the optical anisotropy of an ensemble of GNRs, alignment of the<br />

GNRs is required. Furthermore, oriented GNRs may have controllable end-to-end coupling<br />

between their LSPRs (in contrast to the distribution of relative orientations found in isotropic<br />

dispersions<br />

of GNRs), 30 especially at high GNR loading levels. Several independent reports of the optical<br />

properties of GNRs aligned with electric fields 31 or in stretched polymers 16 have established<br />

that linearly polarized light oriented with the electric field parallel to the GNR long axis<br />

83


excites the LSPR but does not excite the transverse surface plasmon. Here, we show similar<br />

results for aligned GNRs in oriented electrospun PEO microfibers.<br />

Optical absorbance spectra have been collected using the custom-built UV-vis<br />

spectrophotometer described earlier. As a benchmark, spectra acquired at different<br />

polarization angles from GNRs randomly dispersed in a PEO film measuring ~500 �m thick<br />

are presented in Figure 4a and reveal the existence of clearly discernible and angle-<br />

independent LSPR peaks near ~520 and ~850 nm. These signature features are likewise<br />

observed for GNRs dispersed in water, confirming that the bulk, semicrystalline PEO has<br />

little effect on GNR orientation. The absorption spectra of aligned GNRs in electrospun PEO<br />

microfibers measuring 200 nm in diameter (Figure 4b) strongly depend on the polarization<br />

angle. As expected, the LSPR peak at 804 nm is most pronounced when the polarizer is<br />

parallel to n (and to the GNR longitudinal axis) at 0° but vanishes when the polarizer is<br />

perpendicular to n at 90°. Similarly, the absorbance in the 500-600 nm region decreases<br />

when the polarizer is parallel to n at 0° because the transverse surface plasmon is not excited.<br />

The position of the LSPR peak indicates a redshift of 40 nm relative to a random GNR<br />

dispersion in water (cf. Figure S-1 in the Supporting Information). The analogous LSPR peak<br />

in the random dispersion of GNRs in a PEO film is centered at 847 nm (Figure 4a), which<br />

corresponds to a further red shift of ~40 nm. Shifts in the LSPR wavelength may arise from<br />

several sources, such as interparticle coupling, 32 differences between the refractive indices of<br />

PEO and water, and the crystallinity of PEO.<br />

84


The absorbance spectra in Figure 4b have been processed to remove background<br />

contributions from the glass substrate and from optical scattering of the polymer fibers.<br />

Spectra for the GNR-containing and control fiber specimens prior to subtraction are included<br />

for examination in Figure S-2 of the Supporting Information. In contrast to polymer thin<br />

films, polymer fibers contribute a scattering signal that is significantly greater than the<br />

absorbance of the GNRs to extinction, which results in absorbance spectra from the oriented<br />

GNRs that are noisier. All spectra are smoothed using a 17-point Savitzky-Golay numerical<br />

procedure. 33 An aligned PEO microfiber mat without GNRs is selected as an appropriate<br />

control specimen for Figure 4b. The difference in specimen density accompanying the<br />

incorporation of GNRs is taken into account by Beer’s law:<br />

A = ��z, (3)<br />

where A denotes the absorbance, μ is the extinction coefficient and z is specimen thickness.<br />

In Figure 4a, the absorbance spectra of randomly dispersed GNRs in a PEO thin film display<br />

a minimum near 650 nm. It immediately follows that �GNR ≈ �c at this wavelength, thereby<br />

yielding AGNR/zGNR ≈ Ac/zc, where the subscripted c represents the control specimen without<br />

GNRs. The thicknesses of the control and GNR-containing specimens can therefore be<br />

related by their absorbance values at 650 nm, and control spectra have been subtracted from<br />

the spectrum for the GNR-containing fibers. After this correction, the spectra are normalized<br />

to zero and unity. We note, however, that �GNR > �c at 650 nm, because the GNRs have a<br />

small, but non-zero, absorbance at 650 nm. Consequently, the background correction in<br />

Figure 4b does not completely remove spectral contributions originating from polymer fiber<br />

85


scattering, which is responsible for the peak present at ~600-700 nm. The optical anisotropy<br />

of the GNRs could be further increased by improving the alignment of the GNRs within<br />

polymer microfibers and the parallel orientation of the microfibers.<br />

In conclusion, we have demonstrated a scalable method for controlling and maintaining the<br />

nanoscale orientation of GNRs with long-range macroscopic order over a distance of several<br />

centimeters. Here, GNRs with an aspect ratio of 3.1 exhibit excellent alignment with their<br />

longitudinal axes parallel to n for electrospun polymer nano/microfibers with diameters of<br />

40-600 nm, and they maintain substantial alignment in microfibers measuring up to 3000 nm<br />

in diameter. While fiber diameter is found to play a crucial role in GNR alignment, GNR<br />

concentration can be varied with no discernible impact on the net degree of alignment.<br />

Electron diffraction measurements of the aligned GNRs confirm preferred orientation of the<br />

{100} and {111} GNR planes. Optical absorbance spectroscopy measurements performed on<br />

microscopically aligned GNRs in macroscopically aligned electrospun fibers demonstrate<br />

that the LSPR bands are polarization dependent and display maximum absorption when the<br />

polarizer is parallel to n.<br />

Acknowledgement.<br />

This work was supported by the National Science Foundation (CBET-0967559 and a<br />

Graduate Research Fellowship for K. E. R.) and startup funds from North Carolina State<br />

University. We thank Prof. Jesse Jur for helpful discussions and Prof. Benjamin Wiley for<br />

assistance with preliminary optical measurements.<br />

86


Figures<br />

Figure 2.1. TEM image of GNRs deposited from an aqueous suspension onto a carboncoated<br />

TEM grid. The inset shows the distribution of measured aspect ratios of the GNRs,<br />

which measure 49 nm long and 17 nm in diameter on average.<br />

87


Figure 2.2. SEM image of macroscopically-aligned electrospun PEO fibers containing<br />

GNRs.<br />

88


Figure 2.3. Aligned GNRs in electrospun PEO nano/microfibers as functions of fiber<br />

diameter and GNR volume fraction (��: (a) 40 nm and (�� = 0.006, (b) 50 nm and (�� =<br />

0.045, (c) 650 nm and (���= 0.035, and (d) 3000 nm and (��= 0.031. A selected-area<br />

electron diffraction pattern of the corresponding sample is included as an inset in (b).<br />

89


Figure 2.4. Absorbance spectra for (a) randomly oriented GNRs in a PEO film<br />

measuring ~500 �m thick at different polarization angles and (b) GNRs aligned within<br />

electrospun PEO microfibers measuring ~200 nm in diameter at polarization angles varying<br />

from 0° (parallel to the fiber axis n) to 90° (perpendicular to n). In both cases, the data are<br />

color-coded and labeled in (a).<br />

90


References<br />

(1) Tao, A. R.; Habas, S.; Yang, P. D. Small 2008, 4, 310.<br />

(2) Merkel, T. C.; Freeman, B. D.; Spontak, R. J.et al. Science 2002, 296, 519.<br />

(3) Sun, Y. G.; Xia, Y. N. Science 2002, 298, 2176.<br />

(4) Barnes, W. L.; Dereux, A.; Ebbesen, T. W. Nature 2003, 424, 824.<br />

(5) Schön, G.; Simon, U. Colloid Polym. Sci. 1995, 273, 101.<br />

(6) Whitney, T. M.; Jiang, J. S.; Searson, P. C.et al. Science 1993, 261, 1316.<br />

(7) Kelly, K. L.; Coronado, E.; Zhao, L. L.et al. J. Phys. Chem. B 2003, 107, 668.<br />

(8) Link, S.; El-Sayed, M. A. J. Phys. Chem. B 1999, 103, 8410.<br />

(9) Murphy, C. J.; Thompson, L. B.; Chernak, D. J.et al. Curr. Opin. Colloid Interface Sci. 2011,<br />

16, 128.<br />

(10) Jana, N. R.; Gearheart, L.; Murphy, C. J. J. Phys. Chem. B 2001, 105, 4065.<br />

(11) Weissleder, R. Nat. Biotechnol. 2001, 19, 316; Oldenburg, A. L.; Hansen, M. N.; Ralston, T.<br />

S.et al. J. Mater. Chem. 2009, 19, 6407; Durr, N. J.; Larson, T.; Smith, D. K.et al. Nano Lett. 2007, 7,<br />

941.<br />

(12) Nikoobakht, B.; Wang, Z. L.; El-Sayed, M. A. J. Phys. Chem. B 2000, 104, 8635; Sau, T. K.;<br />

Murphy, C. J. Langmuir 2005, 21, 2923.<br />

(13) Vaia, R. A.; Maguire, J. F. Chem. Mater. 2007, 19, 2736.<br />

(14) Sönnichsen, C.; Alivisatos, A. P. Nano Lett. 2005, 5, 301.<br />

(15) Chhetri, R. K.; Kozek, K. A.; Johnston-Peck, A. C.et al. Phys. Rev. E 2011, 83, 040903.<br />

(16) van der Zande, B. M. I.; Pagès, L.; Hikmet, R. A. M.et al. J. Phys. Chem. B 1999, 103, 5761;<br />

Pérez-Juste, J.; Rodríguez-González, B.; Mulvaney, P.et al. Adv. Funct. Mater. 2005, 15, 1065;<br />

91


Murphy, C. J.; Orendorff, C. J. Adv. Mater. 2005, 17, 2173; Li, J. F.; Liu, S. Y.; Liu, Y.et al. Appl.<br />

Phys. Lett. 2010, 96, 263103.<br />

(17) Deshmukh, R. D.; Liu, Y.; Composto, R. J. Nano Lett. 2007, 7, 3662; Nie, Z. H.; Fava, D.;<br />

Kumacheva, E.et al. Nat. Mater. 2007, 6, 609.<br />

(18) Sánchez-Iglesias, A.; Grzelczak, M.; Pérez-Juste, J.et al. Angew. Chem., Int. Ed. 2010, 49,<br />

9985; Correa-Duarte, M. A.; Pérez-Juste, J.; Sánchez-Iglesias, A.et al. Angew. Chem., Int. Ed. 2005,<br />

44, 4375.<br />

(19) Hore, M. J. A.; Composto, R. J. Acs Nano 2010, 4, 6941.<br />

(20) Chang, H. L.; Tian, L.; Abbas, A.et al. Nanotechnology 2011, 22, 275311.<br />

(21) Nikoobakht, B.; El-Sayed, M. A. Chem. Mater. 2003, 15, 1957.<br />

(22) Li, D.; Wang, Y. L.; Xia, Y. N. Nano Lett. 2003, 3, 1167.<br />

(23) Bates, M. A.; Frenkel, D. J. Chem. Phys. 2000, 112, 10034.<br />

(24) Yarin, A. L.; Koombhongse, S.; Reneker, D. H. J. Appl. Phys. 2001, 90, 4836.<br />

(25) Forest, M. G.; Zhou, R. H.; Wang, Q. Int. J. Numer. Anal. Mod. 2007, 4, 478.<br />

(26) Dror, Y.; Salalha, W.; Khalfin, R. L.et al. Langmuir 2003, 19, 7012.<br />

(27) Bashouti, M.; Salalha, W.; Brumer, M.et al. ChemPhysChem 2006, 7, 102.<br />

(28) Johnson, C. J.; Dujardin, E.; Davis, S. A.et al. J. Mater. Chem. 2002, 12, 1765; Petrova, H.;<br />

Perez-Juste, J.; Zhang, Z. Y.et al. J. Mater. Chem. 2006, 16, 3957.<br />

(29) Hermans, J. J.; Hermans, P. H.; Vermaas, D.et al. Recl. Trav. Chim. Pays-Bas-J. Roy. Neth.<br />

Chem. Soc. 1946, 65, 427.<br />

(30) Wang, Y.; DePrince, A. E.; Gray, S. K.et al. J. Phys. Chem. Lett. 2010, 1, 2692.<br />

(31) van der Zande, B. M. I.; Koper, G. J. M.; Lekkerkerker, H. N. W. J. Phys. Chem. B 1999,<br />

103, 5754.<br />

92


(32) Jain, P. K.; Eustis, S.; El-Sayed, M. A. J. Phys. Chem. B 2006, 110, 18243; Vial, S.;<br />

Pastoriza-Santos, I.; Pérez-Juste, J.et al. Langmuir 2007, 23, 4606.<br />

(33) Savitzky, A.; Golay, M. J. E. Anal. Chem. 1964, 36, 1627.<br />

93


Abstract<br />

CHAPTER III<br />

Magnetic Field-Induced Alignment of Nanoparticles in Electrospun<br />

Microfibers<br />

We report on the facile and switchable alignment of superparamagnetic iron oxide<br />

nanoparticles measuring ~18 nm in diameter in electrospun microfibers. Application of a<br />

magnetic field perpendicular to the electric field employed during electrospinning yields<br />

polymeric microfibers with nanoparticles aligned in one-dimensional arrays, thereby<br />

providing control over when and where the nanoparticles align. According to electron<br />

microscopy, the length over which alignment is desired can be judiciously selected, thereby<br />

making these nanomaterials excellent candidates for nanotechnologies requiring nanoscale<br />

alignment on-demand. Concurrent alignment of the electrospun fibers using established<br />

procedures provides a viable route to organic/inorganic materials possessing anisotropic<br />

properties that reflect multiscale alignment.<br />

Introduction<br />

Previous efforts to achieve nanoparticle alignment in polymeric matrices have relied on a<br />

variety of process strategies, such as three-dimensional superlattices, 1 deoxyribonucleic acid 2<br />

or block copolymer 3 templating, surface nanolithography, 4 electrostatic desalting transition, 5<br />

and surface chemical modification. 6 Electrospinning is a rapidly developing fabrication<br />

technique that produces solid polymeric fibers with diameters ranging from several tens of<br />

94


nanometers up to several microns, a high surface-area-to-volume ratio, and the potential for<br />

porosity at multiple length scales. 7 As such, it provides an attractive approach to the<br />

formation of nano/microscale polymeric fibers containing nanoparticles that are spatially<br />

restricted. Further control over nanoparticle positioning can be achieved within such fibers<br />

by post-crystallization of the polymer matrix, 8 the electric field employed during<br />

electrospinning, 9 development of a polymeric nanostructure via self-assembly, 10 or strategic<br />

use of coaxial electrospinning. 11 These methodologies and others, however, depend on the<br />

intrinsic nature of the polymer matrix and/or the processing conditions associated with<br />

electrospinning, and are not intrinsically switchable. Here, we postulate that control over the<br />

location and alignment of nanoparticles within electrospun fibers can be achieved through the<br />

straightforward use of an external electromagnetic field applied during electrospinning. As<br />

with electrospinning, several design issues, such as the strength of the magnet and its<br />

orientation/position with respect to the onset of the polymer jet at the Taylor cone, warrant<br />

consideration. Unlike other approaches intended to align nanoparticles during<br />

electrospinning, an external magnetic field can be applied and removed at any time, thereby<br />

permitting switchable alignment that can even be pulsed. As a consequence of switchable<br />

alignment, the magnetic properties of the nanoparticles can be correspondingly changed from<br />

superparamagnetic to ferromagnetic, 12 which may allow for future read-write capability.<br />

Existing nanotechnologies require aligned nanoparticle arrays in surface-enhanced Raman<br />

scattering (SERS) substrates for use in ultrasensitive analytical tools, 13 micromechanical<br />

95


sensors, 14 nanoscale barcodes, 15 and protein separation. 16 Prior studies have shown that<br />

magnetic fields can be applied during the electrospinning of polymer/nanoparticle systems to<br />

improve alignment of the fibers, but not of the nanoparticles within the fibers. Parallel-<br />

positioned permanent magnets with field strengths ranging from 25 to 120 mT have been<br />

placed, for instance, on the collector plate during the electrospinning of polymer fibers<br />

containing Fe3O4 nanoparticles to align the electrospun fibers. 17 Deflection of entire fibrous<br />

mats generated by electrospinning and containing magnetite nanoparticles in the presence of<br />

a magnetic field has also been observed. 9 In addition, magnetic fields have also been used to<br />

control the spatial arrangement of nanoscale objects within bulk, not electrospun, polymer<br />

matrices. Magneto-polymer nanocomposite particles measuring 200 nm in diameter are, for<br />

instance, reported 18 to align in hydrogel nanocomposites under a magnetic field of 1.5 T.<br />

Even multi-wall carbon nanotubes dispersed in a monomer precursor are found 19 to align<br />

when exposed to an external magnetic field during matrix polymerization. The use of an<br />

external magnetic field has likewise been found to have an effect on biological systems by<br />

organizing cell rods, seeded on magnetically susceptible fiber bundles, into three-<br />

dimensional tissue constructs. 20 In the spirit of these prior observations, the objective of the<br />

present work is to create one-dimensional, periodic arrays of magnetic nanoparticles in<br />

electrospun polymer fibers by applying a magnetic field, in conjunction with an electric field,<br />

during electrospinning.<br />

96


Experimental<br />

ε-Polycaprolactone (PCL) with a molecular weight of 80 kDa according to the<br />

manufacturer was provided by Solvay (Warrington, England). Reagent-grade chloroform for<br />

electrospinning was obtained from Fisher Scientific (Fairlawn, NJ). Iron oxide nanoparticles<br />

with an average diameter of 17.6 nm were prepared by thermal decomposition of iron oleate<br />

in the presence of oleic acid as the capping agent in high boiling hydrocarbons (docosane and<br />

eicosane) according to the procedures described elsewhere. 21,22 X-ray diffraction data<br />

confirmed that as-synthesized nanoparticles are crystalline and contain mostly wüstite (Fe(1-<br />

x)O) and some spinel (most likely Fe3O4), 21 but at the same time, they are superparamagnetic<br />

iron oxide nanoparticles (SPIONs) according to magnetic measurements. 23 The resultant<br />

SPIONs were precipitated with a mixture of acetone and hexane and then separated by<br />

centrifugation and suspended in chloroform at concentrations between 10 and 50 mg/mL.<br />

Suspensions for electrospinning were prepared by first dissolving PCL in chloroform and<br />

then adding appropriate amounts of the SPION suspension to yield 5 wt% PCL at three<br />

different SPION concentrations: 0.5, 1.0 and 2.5 vol%.<br />

The in-house electrospinning unit, operated at 8 kV, employed a syringe pump and an Al<br />

collection target. The separation distance was 12 cm, and the solution flow rate varied from<br />

10 to 35 �L/min. A horseshoe electromagnet (Edmund Scientific, Tonawanda, NY) was<br />

connected to a 6 V battery to generate a 26 mT magnetic field, which corresponds to the most<br />

uniform field as measured by a DC magnetometer (Alphalab Inc., Salt Lake City, UT). As<br />

schematically depicted in Figure 3.1, the magnet was positioned parallel to and below the<br />

97


syringe needle. Electrospun fibers were collected with and without the magnetic field applied<br />

onto the Al plate, as well as onto carbon-coated transmission electron microscopy (TEM)<br />

grids taped to the plate. To discern the spatial distribution of SPIONs within the PCL fibers,<br />

TEM was performed on a field-emission Hitachi HF2000 microscope operated at 200 kV.<br />

The magnetic behavior of aligned 7 fiber mats was examined on a Quantum Design MPMS<br />

superconducting quantum interference device (SQUID) for fibers electrospun from PCL<br />

dissolved in a 30 mg/mL SPION suspension. Magnetization curves were recorded at 300 K<br />

from fiber mats possessing an average mass of 1.5 mg. The magnetizing field strength for<br />

embedded samples ranged from �20 to +20 kOe, whereas saturation was reached at 6 kOe as<br />

the field strength varied from �80 to 80 kOe for unembedded SPIONs.<br />

Results and Discussion<br />

The polymer solution concentration, separation distance and voltage used to electrospin<br />

PCL fibers have been judiciously selected to yield microfibers with diameters ranging from<br />

100 to 500 nm so that their interior morphologies could be interrogated by TEM. Field-<br />

emission scanning electron microscopy (SEM) results (provided in the Supporting<br />

Information) confirm that these PCL microfibers with and without SPIONs possess an<br />

average diameter of ~300 nm. The microfibers exhibit slight dimpling on their surface, and<br />

very few bead defects containing large SPION aggregates (most likely formed by<br />

nanoparticles with insufficient surface functionalization in the suspension reservoir) develop.<br />

In the absence of an applied magnetic field, SPIONs measuring 17.6 nm in diameter appear<br />

to be randomly distributed throughout the PCL microfibers, as seen in the TEM images<br />

98


provided in Figure 3.2 for microfibers differing in SPION concentration: 0.5 vol% (Figure<br />

3.2a) and 2.5 vol% (Figure 3.2b). A TEM image of the SPIONs drop cast from suspension is<br />

included in the inset of Figure 3.2a for reference. Since these SPIONs are coated with oleic<br />

acid which is fairly compatible with PCL, relatively little aggregation is observed in the PCL<br />

microfibers. Few small SPION aggregates, not evident in these images but found in other<br />

specimens, appear as spheroidal clusters with no preferential orientation. Complementary<br />

energy-dispersive spectroscopy analysis (data provided as Supporting Information) reveals<br />

that few, if any, SPIONs reside on the surface of the microfibers.<br />

An electromagnet is selected for the present study since it generates a magnetic field only<br />

when an electric current is flowing, in which case the magnetic field can be immediately<br />

terminated by stopping the current. Since the objective of this study is to create aligned<br />

arrays of magnetic nanoparticles on-demand and since many types of nanoparticles with<br />

various degrees of magnetic susceptibility exist, we have elected to use superparamagnetic<br />

nanoparticles due to their ability to respond to an external magnetic field but remain<br />

nonresponsive when the field is removed. 24 In the field-induced alignment experiments, the<br />

degree to which SPIONs respond to an externally applied magnetic field governs the extent<br />

to which alignment will occur. For a spherical particle, the magnetic moment (mp) is related<br />

to particle size by 25<br />

��<br />

mp � �d 3<br />

mMS<br />

6<br />

99<br />

(1)


where dm is the particle diameter and MS denotes the saturation magnetization. On the basis<br />

of Eq. 1, very small nanoparticles may have an insufficient magnetic moment to respond to<br />

the applied magnetic field and undergo noticeable alignment. When an external magnetic<br />

field is applied, however, attractive dipolar interactions arise between adjacent SPIONs.<br />

Aggregates of SPIONs develop to maximize their magnetic moment in response to the<br />

magnetic field and align along the magnetic field lines created by the electromagnet. Under<br />

these conditions, the anisotropic field-induced magnetic dipolar interactions promote the<br />

formation of highly elongated aggregates (arrays), rather than random distributions, of<br />

SPIONs.<br />

The syringe portrayed in Figure 3.1 is positioned within several centimeters of the open<br />

end of the horseshoe electromagnet (where the magnetic field lines are the most<br />

concentrated). The optimal distance is ~1.5 cm, which suggests that this position identifies<br />

where the magnetic field is the most uniform. The distance between the poles of the magnet<br />

is 2 cm, with one pole placed directly under the Taylor cone (reasons for which are discussed<br />

later) and the other located 2 cm closer to the collection plate. Although the horseshoe shape<br />

of the electromagnet increases the strength of the magnet, we recognize that the field is not<br />

uniform. This issue is currently being addressed by using electromagnets varying in physical<br />

shape and field uniformity. Another factor influencing the ability of the SPIONs to align in<br />

electrospun PCL microfibers is their concentration in the suspension reservoir. The attractive<br />

forces between two SPIONs whose dipole moments are aligned must be sufficient to<br />

overcome the matrix viscosity for particle alignment to ensue, and the magnitude of such<br />

100


forces depends on the interparticle distance, which relates to the particle concentration. 26 A<br />

sensitivity analysis for 17.6 nm SPIONs reveals that a minimum SPION concentration of ca.<br />

0.05 vol% is required for discernible alignment in the present study. At higher<br />

concentrations, the interparticle distance decreases and greater attraction, or chaining,<br />

between neighboring SPIONs occurs. The maximum SPION concentration is set by other<br />

considerations, namely, suspension conductivity and viscosity, as well as undesirable SPION<br />

aggregation.<br />

Examples of aligned SPIONs measuring 17.6 nm in diameter in electrospun PCL<br />

microfibers at a concentration of 0.5 vol% are presented in Figure 3.3. In all cases, the<br />

electromagnet is connected first, followed by the high-voltage power supply for<br />

electrospinning. The 6 V battery remains outside the electric field to prevent the possibility<br />

of a short circuit. In images such as the one displayed in Figure 3.3a, the SPIONS remain<br />

aligned in one-dimensional arrays parallel to the fiber surface over lengths exceeding 1.5 �m<br />

insofar as the electromagnet remains active. These long arrays are visible in most microfibers<br />

of a mat collected on a TEM grid, and they appear on multiple grids used to collect the same<br />

sample. In Figure 3.3b (including the inset), the effect of the magnetic field is not constant<br />

(most likely due to the reasons discussed below), which results in shortened, but nonetheless<br />

highly aligned, SPION arrays of variable thickness. Although such arrays constitute the<br />

predominant feature when the magnetic field is applied, some randomly distributed SPIONs<br />

remain. Fine adjustment of both electrospinning conditions and solution properties has been<br />

unable to resolve this lack of position uniformity within the microfibers. Possible reasons for<br />

101


this shortcoming are three-fold: (i) the magnetic field produced by the horseshoe-shaped<br />

magnet is inherently not uniform; (ii) instabilities in the jet, 27 coupled with fiber whipping<br />

that occurs between the Taylor cone and the collection plate; and (iii) nanoparticles residing<br />

in the Taylor cone, as previously reported. 1 All of these considerations could result in fibers<br />

that vary in position with respect to the applied magnetic field. In both cases, the strength of<br />

the magnetic field, and thus the uniformity and contiguity of SPION alignment,<br />

simultaneously vary. One way to enhance the spatial characteristics of SPION arrays is by<br />

aligning the microfibers between electrodes instead of generating a random mat of<br />

microfibers on a collection plate. 28 However achieved, improved control over the position of<br />

the fiber within the field is ultimately expected to promote uniform SPION alignment that<br />

can be pulsed or otherwise induced on-demand.<br />

Unlike ferromagnetic materials, magnetization of the superparamagnetic SPIONs under<br />

investigation only occurs under an external magnetic field. When this field is removed, the<br />

net magnetization of the SPION dispersion becomes close to zero. The magnetization (M)<br />

behavior of SPIONs measuring 17.6 nm in diameter is presented as a function of magnetizing<br />

field strength (H) at 300 K in the SQUID hysteresis plot shown in Figure 3.4a. The value of<br />

MS for the unembedded SPIONs is 38.8 emu/g, whereas that for SPIONs encapsulated within<br />

electrospun PCL microfibers is sharply lowered to 0.77�0.94 emu/g. This nontrivial<br />

reduction in MS for the SPIONs residing in electrospun PCL microfibers with and without an<br />

external magnetic field is due largely to the decreased mobility of the SPIONs, a result that<br />

has been previously reported for embedded nanoparticles. 29 If the matrix viscosity is<br />

102


sufficiently high, only the magnetic moment of a SPION can rotate in the presence of an<br />

external field, but the SPION itself may be unable to do so. In light of these results, the<br />

applied magnetic field must be positioned in the vicinity of the Taylor cone formed during<br />

electrospinning (where solvent remains and the matrix viscosity is relatively low). This<br />

location has been selected here to maximize the ability of the external magnetic field to<br />

induce SPION alignment in electrospun PCL microfibers.<br />

According to Figure 3.4, the SPIONs alone and embedded in PCL exhibit a<br />

superparamagnetic loop characterized by a minimum coercivity and a remanent<br />

magnetization at room temperature. The mean magnetic moment (�) and MS are extracted<br />

from each hysteresis curve by fitting the data to the Langevin function, 30 viz.,<br />

M � MS coth �H �� ��<br />

�� ���<br />

��k<br />

BT��<br />

k ��<br />

��<br />

��<br />

BT<br />

�� (2)<br />

�� �H ��<br />

where kB is the Boltzmann constant and T denotes absolute temperature. Regressed values of<br />

��<br />

MS and � for unembedded SPIONs (Figure 3.4a) are 38.8 emu/g and 3.37 10 �17 emu,<br />

respectively. Similarly, from Figure 3.4b, the values of MS and � determined for randomly<br />

dispersed SPIONs are 0.94 emu/g and 1.26 10 �16 emu, respectively, whereas those for<br />

magnetically-aligned SPIONs are 0.77 emu/g and 1.37 10 �16 emu, respectively. The inset<br />

in Figure 3.4a displays magnetization hysteresis curves recorded at relatively low fields and<br />

reveals that the differences in coercivity and remanent magnetization are, for the most part,<br />

negligible at ambient temperature, which is important for applications conducted at ambient<br />

103


conditions. Such differences are, however, expected to be more apparent at low temperatures.<br />

In fact, previous studies 31 conducted at low temperatures (ca. 5 K) report that arrays of<br />

SPIONs possess a noticeably higher remanence and coercivity than randomly distributed<br />

nanoparticles. No such relation is evident from our results at ambient temperature. This may<br />

be due to the fact that the interparticle distance in our systems is ~5 nm, in which case little,<br />

if any, magnetostatic coupling occurs. 26 Thus, even though the SPIONs are arranged in<br />

unidirectional arrays, their magnetic behavior is still consistent with individual particles<br />

rather than a connected nanowire. It must be recognized, however, that the values of MS are<br />

mass-normalized by multiplying the mass ratio of SPION:PCL in the suspension prior to<br />

electrospinning by the mass of the fiber mat (~1.5 mg). Therefore, the corresponding MS<br />

values are sensitive to the SPION populations present. With the possibility of aggregated<br />

SPION clusters forming in suspension or during electrospinning, the calculated SPION mass<br />

in a given specimen may not be sufficiently accurate to permit direct comparison. In contrast,<br />

values of � are not sensitive to the mass present and reveal an interesting trend that is<br />

amenable to computational modeling: the mean magnetic moment increases with increasing<br />

SPION alignment.<br />

Conclusions<br />

Nanoscale alignment of SPIONs with a mean diameter of 17.6 nm has been achieved<br />

through the use of magnetic field-assisted electrospinning. In this case, an electromagnetic<br />

field of 26 mT is applied perpendicular to the electric field required for electrospinning,<br />

resulting in microfibers that contain one-dimensional arrays of SPIONs. In the present<br />

104


system investigated, a SPION concentration of greater than 0.05 vol% is required to induce<br />

discernible alignment. The arrays thus produced can extend beyond 1 �m under a constant<br />

magnetic field, or they can be controllably shortened by pulsing the magnetic field. The latter<br />

finding suggests that it may be possible in the future to write information to individual<br />

microfibers on the basis of array lengths. Moreover, alignment of the electrospun microfibers<br />

through the use of various established methods can yield nanocomposites with multiscale<br />

(i.e., nanoscale and macroscale) anisotropic properties. Complementary SQUID<br />

measurements at ambient temperature reveal that the saturation magnetization is significantly<br />

lower for SPIONs in electrospun fibers with or without magnetic field-induced alignment<br />

than for unembedded SPIONs. Improving alignment, however, appears to increase the mean<br />

magnetic moment, which provides evidence that nanoscale alignment of SPIONs affects their<br />

intrinsic physical properties.<br />

Acknowledgments<br />

This work was supported by a National Science Foundation Graduate Research<br />

Fellowship (K. E. R.), a North Carolina Space Grant, and the NATO Science for Peace<br />

Program (L. M. B) and Indiana University Faculty Research Program (L. M. B.). We thank<br />

Professor Frank Tsui (University of North Carolina at Chapel Hill) for use of his SQUID, and<br />

Mr. Matt Wolboldt for insightful discussions<br />

105


Figures<br />

Figure 3.1. Schematic illustration of the magnetic field-assisted electrospinning setup<br />

used in this study. Note the position of the electromagnet, which yields a magnetic field that<br />

is perpendicular to the electric field employed during electrospinning.<br />

106


Figure 3.2. TEM images of randomly dispersed SPIONs in electrospun PCL<br />

microfibers varying in SPION concentration (in vol%): (a) 0.5 and (b) 2.5. A TEM image of<br />

SPIONs measuring 17.6 nm in diameter and drop cast from chloroform is included in the<br />

inset of (a). The scalemarker in the inset corresponds to 50 nm.<br />

107


Figure 3.3. TEM images of magnetic field-aligned SPIONs, measuring 17.6 nm in<br />

diameter, in PCL microfibers illustrating long, contiguous arrays in (a), and shorter, pulsed<br />

arrays in (b). The scalemarker in the inset corresponds to 100 nm.<br />

108


Figure 3.4. Magnetization (M) hysteresis curves at 300 K as a function of the<br />

magnetizing field strength (H) for SPIONs measuring 17.6 nm in diameter. In (a), the<br />

hysteresis curves are measured from unembedded SPIONs ( ), as well as randomly<br />

dispersed and magnetically aligned SPIONs in electrospun PCL microfibers (blue and red,<br />

respectively). The inset shows magnetization hysteresis curves recorded for the embedded<br />

SPIONS at low fields and ambient temperature. In (b), the hysteresis curves from the<br />

SPIONs embedded in PCL microfibers (see the corresponding diagrams) are fitted to Eq. 2 in<br />

the text (solid lines) to discern the saturation magnetization and mean dipole moment from<br />

each dataset.<br />

109


Supporting Information<br />

Samples for examination by field-emission scanning electron microscopy (SEM) were<br />

prepared by electrospinning fiber mats directly onto silicon wafers that could be inserted into<br />

the specimen chamber of a JEOL 6400F field-emission microscope, which was operated at<br />

an accelerating voltage of 5 kV. To reduce charging effects, the fibers were sputter-coated<br />

with a thin conductive layer (~6 nm) of Au/Pd at ambient temperature. Energy-dispersive<br />

spectroscopy (EDS) was performed with an Oxford Isis x-ray detector mounted in a Hitachi<br />

S3200 variable-pressure electron microscope and equipped with an ultrathin window for light<br />

element analysis.<br />

An example of a PCL fiber mat containing SPIONs is presented in Figure 3.S1. While the<br />

fibers vary considerably in size, the average diameter of fibers acquired under different<br />

experimental conditions is about 300 nm. Some of the fibers appear to possess a dimpled<br />

surface, which is evident in the inset (a large fiber is selected for display to facilitate<br />

scrutinization of the fiber surface). A representative EDS spectrum, as well as an<br />

enlargement of a portion thereof, of a PCL fiber (with 2.5 wt% SPIONs measuring 17.6 nm<br />

in diameter) subjected to a magnetic field during electrospinning is provided in Figure 3.S2.<br />

Elements detected in this and related spectra include C from the fibers, Si and O from the<br />

silicon wafer, and Au and Pd from the conductive coating. Peaks corresponding to the K�<br />

and L lines of Fe (in the vicinities of 6.4 and 0.7 keV, respectively) are absent, indicating that<br />

the SPIONS are, for the most part, embedded within the fibers and do not reside, to an<br />

appreciable extent, on the fiber surface.<br />

110


Figure 3.S1. SEM image of SPION-containing PCL fibers electrospun in the presence of<br />

an external magnetic field. The inset shows evidence of surface dimpling on a large fiber.<br />

The scalemarker in the inset corresponds to 2 �m.<br />

111


Figure 3.S2. EDS spectrum of a SPION-containing PCL fiber electrospun in the<br />

presence of an external magnetic field. The elements responsible for the observed peaks are<br />

labeled, and the x-ray energies associated with the K� and L lines of Fe are identified by the<br />

blue lines.<br />

112


References<br />

(1) Miyauchi, M.; Simmons, T. J.; Miao, J. J.; Gagner, J. E.; Shriver, Z. H.; Aich, U.; Dordick, J.<br />

S.; Linhardt, R. J., ACS Appl. Mater. Interfaces 2011, 3, 1958-1964.<br />

(2) Mucic, R. C.; Storhoff, J. J.; Mirkin, C. A.; Letsinger, R. L., J. Am. Chem. Soc. 1998, 120,<br />

12674-12675.<br />

(3) Harris, L. A.; Goff, J. D.; Carmichael, A. Y.; Riffle, J. S.; Harburn, J. J.; St Pierre, T. G.;<br />

Saunders, M., Chem. Mater. 2003, 15, 1367-1377.<br />

(4) Tsirlin, T.; Zhu, J.; Grunes, J.; Somorjai, G. A., Top. Catal. 2002, 19, 165-170.<br />

(5) Yan, M.; Fresnais, J.; Sekar, S.; Chapel, J. P.; Berret, J. F., ACS Appl. Mater. Interfaces<br />

2011, 3, 1049-1054.<br />

(6) Xiong, Y.; Chen, Q. W.; Tao, N.; Ye, J.; Tang, Y.; Feng, J. S.; Gu, X. Y., Nanotechnology<br />

2007, 18.<br />

(7) Li, W. J.; Laurencin, C. T.; Caterson, E. J.; Tuan, R. S.; Ko, F. K., Journal of Biomedical<br />

Materials Research 2002, 60, 613-621.<br />

(8) Reneker, D. H.; Chun, I., Nanotechnology 1996, 7, 216-223.<br />

(9) Kim, G. M.; Wutzler, A.; Radusch, H. J.; Michler, G. H.; Simon, P.; Sperling, R. A.; Parak,<br />

W. J., Chem. Mater. 2005, 17, 4949-4957.<br />

(10) Wang, A.; Singh, H.; Hatton, T. A.; Rutledge, G. C., Polymer 2004, 45, 5505-5514.<br />

(11) Wang, B. B.; Li, B.; Xiong, J.; Li, C. Y., Macromolecules 2008, 41, 9516-9521.<br />

(12) Song, T.; Zhang, Y. Z.; Zhou, T. J.; Lim, C. T.; Ramakrishna, S.; Liu, B., Chem. Phys. Lett.<br />

2005, 415, 317-322.<br />

(13) Dyab, A. K. F.; Ozmen, M.; Ersoz, M.; Paunov, V. N., J. Mater. Chem. 2009, 19, 3475-3481.<br />

113


(14) Peng, P.; Chen, Y. Z.; Gao, Y. F.; Yu, J.; Guo, Z. X., Journal of Polymer Science Part B-<br />

Polymer Physics 2009, 47, 1853-1859.<br />

(15) Goubault, C.; Jop, P.; Fermigier, M.; Baudry, J.; Bertrand, E.; Bibette, J., Phys. Rev. Lett.<br />

2003, 91.<br />

(16) Nam, J. M.; Park, S. J.; Mirkin, C. A., J. Am. Chem. Soc. 2002, 124, 3820-3821.<br />

(17) Doyle, P. S.; Bibette, J.; Bancaud, A.; Viovy, J. L., Science 2002, 295, 2237-2237.<br />

(18) Yang, D. Y.; Lu, B.; Zhao, Y.; Jiang, X. Y., Adv. Mater. 2007, 19, 3702-3706.<br />

(19) Nunes, J.; Herlihy, K. P.; Mair, L.; Superfine, R.; DeSimone, J. M., Nano Lett. 2010, 10,<br />

1113-1119.<br />

(20) Kimura, T.; Ago, H.; Tobita, M.; Ohshima, S.; Kyotani, M.; Yumura, M., Adv. Mater. 2002,<br />

14, 1380-1383.<br />

(21) Lee, W. Y.; Cheng, W. Y.; Yeh, Y. C.; Lai, C. H.; Hwang, S. M.; Hsiao, C. W.; Huang, C.<br />

W.; Chen, M. C.; Sung, H. W., Tissue Engineering Part C-Methods 2011, 17, 651-661.<br />

(22) Bronstein, L. M.; Huang, X.; Retrum, J.; Schmucker, A.; Pink, M.; Stein, B. D.; Dragnea, B.,<br />

Chem. Mater. 2007, 19, 3624-3632.<br />

(23) Bronstein, L. M.; Atkinson, J. E.; Malyutin, A. G.; Kidwai, F.; Stein, B. D.; Morgan, D. G.;<br />

Perry, J. M.; Karty, J. A., Langmuir 2011, 27, 3044-3050.<br />

(24) Shtykova, E. V.; Huang, X.; Remmes, N.; Baxter, D.; Stein, B. D.; Dragnea, B.; Svergun, D.<br />

I.; Bronstein, L. M., J. Phys. Chem. B 2007, 111, 18078-18086.<br />

(25) Gupta, A. K.; Gupta, M., Biomaterials 2005, 26, 3995-4021.<br />

(26) Jiles, D., Introduction to magnetism and magnetic materials. Chapman and Hall: Boca Raton,<br />

1998; Vol. 2.<br />

(27) Russier, V.; Petit, C.; Pileni, M. P., J. Appl. Phys. 2003, 93, 10001-10010.<br />

114


(28) Reneker, D. H.; Yarin, A. L.; Fong, H.; Koombhongse, S., J. Appl. Phys. 2000, 87, 4531-<br />

4547.<br />

(29) Li, D.; Wang, Y. L.; Xia, Y. N., Nano Lett. 2003, 3, 1167-1171.<br />

(30) Zhang, D.; Karki, A. B.; Rutman, D.; Young, D. R.; Wang, A.; Cocke, D.; Ho, T. H.; Guo, Z.<br />

H., Polymer 2009, 50, 4189-4198.<br />

(31) Bayat, M.; Yang, H.; Ko, F., Polymer 2011, 52, 1645-1653.<br />

(32) Zhu, J. H.; Wei, S. Y.; Chen, X. L.; Karki, A. B.; Rutman, D.; Young, D. P.; Guo, Z. H., J.<br />

Phys. Chem. C 2010, 114, 8844-8850.<br />

(33) Langevin, P., Annales De Chimie Et De Physique 1905, 5, 70-127.<br />

(34) Park, J. I.; Jun, Y. W.; Choi, J. S.; Cheon, J., Chem. Commun. 2007, 5001-5003.<br />

(35) Petit, C.; Russier, V.; Pileni, M. P., J. Phys. Chem. B 2003, 107, 10333-10336.<br />

115


CHAPTER IV<br />

Using Polymer Blend Morphology to Position Ligand-Functionalized<br />

Abstract<br />

Nanoparticles in Electrospun Polymer Microfibers<br />

Blends of hydrophobic and hydrophilic polymers have been prepared to discern the<br />

feasibility of controlling the spatial location of superparamagnetic iron oxide nanoparticles<br />

(SPIONs) within electrospun microfibers on the basis of thermodynamic compatibility.<br />

Although inorganic nanoparticles tend to aggregate, a carefully designed polymer matrix can<br />

be used to position nanoparticles and achieve unique optical, electronic and/or magnetic<br />

properties. In the present study, poly(ethylene oxide) (PEO) and poly(2-vinyl pyridine)<br />

(P2VP) phase-separate in electrospun microfibers and form discrete, but diffuse, dispersions<br />

at low concentrations of the minority component. At higher concentrations, a core-sheath<br />

structure naturally develops wherein the hydrophobic SPIONs are sequestered, according to<br />

electron microscopy. X-ray diffractometry confirms that the (120) reflection increases in<br />

intensity after SPIONs are added, suggesting that the PEO chains become increasingly<br />

aligned along the fiber axis and that the nanoparticles reside at the PEO/P2VP interface affect<br />

PEO crystallization. We thus achieve control over SPION positioning within these<br />

microfibers without requiring post-modification.<br />

116


Introduction<br />

The formation of polymer nanocomposites has become a preferred means by which to<br />

combine the highly desirable electrical, magnetic, optical and/or mechanical properties of<br />

metals and metal oxides with the flexibility, light weight and processability of inexpensive<br />

and tough polymers. 1 Incorporation of stimuli-responsive nanoscale objects into<br />

homogeneous polymer matrices generally yields functional materials that are suitable for<br />

various applications such as data storage, conductive nanowires, nonwoven sensors,<br />

magnetic filters, and nanoactuation. 2 Metallic nanoparticles, in particular, tend to aggregate<br />

and reduce the population of surface atoms because of the driving force to decrease their total<br />

surface energy. 3 Harnessing control over the spatial location of nanoparticles within a<br />

polymer matrix may enable favorable attributes 4 associated with their electronic, optical,<br />

magnetic, and catalytic properties. One way by which to attain this objective is to employ a<br />

heterogeneous polymer matrix composed of either a macrophase-separated polymer blend 5 or<br />

a microphase-separated block copolymer. 6 Independent studies have demonstrated 7 that<br />

judicious manipulation of enthalpic and entropic contributions to the system free energy by<br />

systematic variation in nanoparticle size, concentration and surface chemistry can be<br />

exploited to position nanoparticles within (micro)phase domains or along interfaces on the<br />

basis of thermodynamic considerations.<br />

Another, process-oriented, means by which to control nanoparticle positioning is<br />

through physical confinement, 8 which can be readily achieved by spin-casting or<br />

electrospinning. Electrospinning is an emerging fabrication technique that produces solid<br />

117


fibers with diameters ranging from several tens of nanometers up to several microns, a high<br />

surface-area-to-volume ratio and high porosity from both polymer solutions and melts. 9 It is<br />

an appealing fiber-spinning strategy due to the relatively straightforward setup and its ability<br />

to tune fiber properties on the basis of both processing and solution variables. A polymer<br />

solution or melt is slowly ejected from a spinneret which is connected to a high voltage<br />

source. As a voltage is applied, a charge develops on the surface of the liquid due to mutual<br />

charge repulsion 10 and as these electrostatic forces exceed the surface tension of the polymer<br />

solution or melt, a charged jet is emitted. This jet has a conical structure defined as a “Taylor<br />

cone” 11 and undergoes a whipping process, 12 during which time the solvent evaporates and is<br />

collected as a dry, randomly oriented polymer fiber mat on the collection plate.<br />

Since the fiber-forming polymer often lacks functionality that is desired for growing<br />

multifunctional applications, functionality can be imparted by several different routes, such<br />

as addition of a second polymer, 13 spinning a block copolymer system, in situ growth of core<br />

or surface functionalities through the incorporation of solution or melt precursors, 14 co- 15 and<br />

multi-axial 16 spinnerets where multiple layers of a fiber can be created, or post-electrospun<br />

surface modification 17 of the fiber surface.<br />

Both miscible and immiscible polymer blends can be electrospun to create new<br />

morphologies of phase-separated nano/microfibers. 18 When two polymers phase-separate<br />

into discrete domains, one component can be selectively removed to yield nanoporous fibers,<br />

which has been reported for polymer pairs such as poly(lactic acid)/poly(vinylpyrrolidone) 19 ,<br />

poly(ethylene oxide)/silk, 20 poly(ethylene oxide)/poly(methyl methacrylate) 21 and<br />

118


poly(glycolic acid)/chitin. 22 When phase separation yields more contiguous biphasic<br />

morphologies, core-sheath structures 23 can develop. A core-sheath microstructure is highly<br />

desirable since it is possible to combine polymers with two vastly different property sets<br />

(e.g., the sheath can be insulating while the core can be conductive 23 ). In addition, the core<br />

may also be selectively removed to form nano/microtubes, 24 which would be beneficial for<br />

time-controlled drug release 25 and sensory applications 26 . Core-sheath fiber microstructures<br />

are formed by several different protocols: careful solvent selection, 27 thermostatic/kinetic<br />

polymer contrast, 28 co-axial electrospinning, 29 or electric field-induced. 21 Although these<br />

specialized procedures have been developed to generate core-sheath microstructures, we<br />

hasten to point out that such microstructures rarely form naturally. Using polymer blends to<br />

spontaneously form core-sheath microstructures during electrospinning is particularly<br />

appealing since the chemistry, molecular weight and concentration of the polymer<br />

constituents can be independently adjusted.<br />

The primary objective of this work is to generate nanocomposite nano/microfibers<br />

containing spatially positioned superparamagnetic iron oxide nanoparticles (SPIONs) by<br />

electrospinning. Control over nanoparticle positioning has been previously achieved within<br />

such fibers by post-crystallization of the polymer matrix, 30 alteration of the electric field<br />

employed during electrospinning, 31 development of a polymeric nanostructure via self-<br />

assembly, 32 or strategic use of coaxial electrospinning. 33 Here, we utilize a blend of<br />

hydrophobic and hydrophilic polymers to discern the feasibility of controlling the spatial<br />

location of SPIONS within electrospun fibers on the basis of thermodynamic compatibility.<br />

119


Experimental<br />

Materials<br />

Poly(2-vinyl pyridine) (P2VP) with a molecular weight of 200 kDa according to the<br />

manufacturer was obtained from Scientific Polymer Products (Ontario, NY), and<br />

poly(ethylene oxide) (PEO) with a molecular weight of 200 kDa according to the<br />

manufacturer was provided by Sigma-Aldrich (Milwaukee, WI). Reagent-grade chloroform<br />

for electrospinning was purchased from Fisher Scientific (Fairlawn, NJ) and used as-<br />

received. Iron oleate, oleic acid and the various solvents used in the nanoparticle synthesis<br />

were acquired from TCI America (Portland, OR) and also used without further purification.<br />

Preparation of Nanoparticles<br />

The SPIONs with diameters ranging from 12 to 24 nm were prepared by thermal<br />

decomposition of iron oleate in the presence of oleic acid as the capping agent in high boiling<br />

hydrocarbons (docosane and eicosane), according to the procedures described elsewhere. 34,35<br />

Previous X-ray diffractometry (XRD) data confirmed 34 that these as-synthesized SPIONs are<br />

crystalline and contain mostly wüstite (Fe(1-x)O) and some spinel (most likely Fe3O4).<br />

According to magnetometry measurements, 36 they are superparamagnetic. The resultant<br />

SPIONs were precipitated with a mixture of acetone and hexane, separated by centrifugation<br />

and suspended in chloroform at concentrations between 10 and 50 mg/mL (Figure 4.1).<br />

120


Preparation of Nano/microfibers<br />

Suspensions for electrospinning were prepared by first dissolving predetermined<br />

quantities of PEO and P2VP (ranging from 0 to 100 wt% PEO) in chloroform and then<br />

adding a target mass of a SPION suspension to yield polymer concentrations ranging from 2<br />

to 8 wt% at three different SPION concentrations: 0.5, 1.0 and 2.5 vol%. The in-house<br />

electrospinning unit, operated at 8 kV, employed a syringe pump operated at flow rates<br />

between 10 and 35 �L/min, and an Al collection target maintained at a separation distance of<br />

12 cm. Electrospun nano/microfibers were collected as unoriented nonwoven mats on the Al<br />

plate, as well as on carbon-coated transmission electron microscopy (TEM) grids adhered to<br />

the plate.<br />

Characterization of Nano/microfibers<br />

To discern the blend morphologies of the electrospun nano/microfibers and the spatial<br />

distribution of SPIONs contained therein, TEM was performed on a field-emission Hitachi<br />

HF2000 microscope operated at 200 kV. To enhance the contrast between the two polymer<br />

phases, the P2VP has been selectively stained with the vapor of 0.5% OsO4(aq) for 7 min.<br />

Corresponding scanning electron microscopy (SEM) images were collected from specimens<br />

sputter-coated with 6 nm of Au/Pd on a JEOL 6400F field-emission microscope operated at 5<br />

kV. Complementary XRD was performed at 2θ angles ranging between 5 and 30° in 0.01°<br />

increments on a Rigaku Smartlab diffractometer with CuKα radiation at a wavelength (�) of<br />

0.1541 nm. Zero-shear viscosity measurements were performed on a TA Instruments AR-G2<br />

121


heometer over shear stresses extending from 0.05 to 50 Pa on 40 mm 2° steel cone plates in<br />

soft bearing mode.<br />

Results and Discussion<br />

Blends of hydrophobic and hydrophilic polymers have been prepared and electrospun to<br />

discern the feasibility of controlling the spatial location of SPIONs within electrospun<br />

nano/microfibers due to preferential compatibility. Both P2VP and PEO are specifically<br />

chosen for this purpose and likewise satisfy two additional criteria: they are of comparable<br />

molecular weight and they are both soluble in chloroform, which is used for electrospinning.<br />

The blends range in concentration from 0 to 100 wt% PEO. Due to differences in mass<br />

density (1.15 g/cm 3 for P2VP and 1.21 g/cm 3 for PEO), the concentration of each polymer<br />

solution is appropriately adjusted to yield electrospun nano/microfibers of comparable<br />

diameter (~1200 nm), thereby eliminating fiber size from consideration in morphological<br />

analyses. Fibers of this size will hereafter be referred to as microfibers. Nonpolar solubility<br />

parameters for PEO, P2VP and chloroform are reported 37 as 15.6, 21.7 and 17.8 MPa 1/2 ,<br />

respectively. On the basis of these values alone, it follows that PEO is more compatible with<br />

chloroform and will remain solvated longer during electrospinning than P2VP. In the event<br />

that a core-sheath microstructure developed, one could reasonably presume that P2VP should<br />

form the core and PEO should form the surrounding sheath (and crystallize).<br />

Field-emission SEM confirms that these PEO/P2VP microfibers with and without<br />

SPIONs possess an average diameter of ~1200 nm and relatively smooth surfaces with very<br />

few bead defects, as evidenced by the images displayed in Figures 4.2a and 4.2b. It is<br />

122


interesting that the defects tend to contain large SPION aggregates (cf. the inset of Figure<br />

4.2b), which most likely formed in the suspension reservoir by nanoparticles with insufficient<br />

surface functionalization. Examples of TEM images acquired from microfibers varying in<br />

composition without SPIONs establish that the PEO phase is indiscernible at low<br />

concentrations (suggesting that the two homopolymers may actually be at least partially<br />

miscible under the conditions employed). Unlike previous studies that have reported 38<br />

nanoparticle-induced changes in nanoscale morphologies, these morphologies are largely<br />

retained when SPIONs at a concentration of 2.5 vol% are added, as in Figure 4.3. Since the<br />

polymers are phase-separated at some blend compositions, the hydrophobic nanoparticles can<br />

be anticipated to preferentially reside in one phase. According to the TEM images of PEO-<br />

rich blends presented in Figure 4.4, the SPIONs, measuring 18 nm in diameter, are largely<br />

sequestered within the P2VP phase, which has been selectively stained and thus appears<br />

more electron-dense (dark). The SPIONs do not adopt a particular alignment and appear to<br />

be randomly distributed throughout the P2VP phase. Since these SPIONs are coated with<br />

oleic acid, which induces mutually hydrophobic interactions between alkyl chains, 39<br />

relatively little nanoparticle aggregation is observed in these microfibers, regardless of blend<br />

composition. Very few small SPION aggregates, not evident in the images shown in Figure<br />

4.3 but occasionally observed in other specimens, appear as spheroidal clusters. The diameter<br />

and concentration of SPIONs has been systematically varied from 12 to 24 nm and from 0.5<br />

to 2.5 vol%, respectively, with no significant effect on their host phase or the accompanying<br />

123


lend morphology. For these reasons, we only present images of SPIONs measuring 18 nm<br />

in diameter at a concentration of 2.5 vol%.<br />

The phase morphology of polymer blends subjected to nonequilibrium processing such as<br />

electrospinning depends sensitively on thermodynamic (e.g., interfacial tension, solubility<br />

and molecular weight) and kinetic (viscoelasticity and molecular mobility) factors. 40<br />

Nanostructure formation within a polymer fiber during electrospinning is estimated to be on<br />

a time scale of ~10 ms due to concurrent solvent evaporation and jet elongation. 41 In the<br />

present PEO/P2VP blends, thermodynamic driving forces seemingly dominate, especially at<br />

intermediate to high concentrations of PEO, even though P2VP in chloroform possesses a<br />

lower viscosity than PEO (discussed later). Such incompatibility is evident from the<br />

differences in solubility parameters described earlier. At low PEO concentrations (e.g., 20<br />

wt% PEO), the absence of two distinct phases is suggestive of phase mixing. One must<br />

exercise caution in drawing this conclusion since the stained P2VP is more electron-dense<br />

than PEO, in which case discrete PEO domains may not be visible within the P2VP matrix.<br />

Analysis of images such as Figure 4.3a confirm that there is no significant change in optical<br />

density along the length of the microfibers. As the concentration of PEO is increased to 30<br />

wt% (Figure 4.3b), seemingly unconnected PEO domains develop along the edge of the<br />

microfibers, thereby forming the onset of core-sheath microstructure. At this blend<br />

composition, the low viscosity of the blend in chloroform allows PEO chains to migrate to<br />

the microfiber-air interface as the chloroform evaporates. In response, the P2VP chains<br />

remain along and thus form the core.<br />

124


To quantify the extent to which the PEO and P2VP are segregated, the interfacial<br />

thickness (�) corresponding to the change in optical density across the interfacial normal has<br />

been measured from as many as 10 different regions in up to 3 different images. We<br />

recognize that these values are only estimates, as the interfacial curvature has not been<br />

considered, but they can provide insight into the incompatibility between PEO and P2VP<br />

upon electrospinning. At 30 wt% PEO, � is approximately 32 nm, which corresponds to a<br />

very diffuse interface. 42 As the concentration of PEO is increased to 40 wt%, P2VP forms the<br />

major phase with small pockets of PEO along the fiber surface. It is interesting to note that<br />

these PEO domains, which range from 175 to 400 nm in length, are oriented perpendicular to<br />

the fiber spinning direction. This orientation may be a consequence of solvent evaporation as<br />

the microfiber solidifies. Although the morphology is somewhat different from what is seen<br />

at the lower PEO concentration, � is comparable at ~29 nm, confirming that the degree of<br />

phase segregation is similar. When the concentration of PEO is further increased to 50 wt%<br />

(Figure 4.3c), two sizes of P2VP domains reside within an apparent PEO matrix. The larger<br />

domain sizes measure ~200 nm across, whereas the smaller ones are ~60 nm, and � increases<br />

noticeably to ~52 nm, which is indicative of enhanced phase mixing. At 60 wt% PEO (Figure<br />

4.3d), the PEO possesses sufficient mobility to migrate in large part to the microfiber-air<br />

interface and � ≈ 25 nm is similar to what it was at lower PEO concentrations. In the case of<br />

the fractal-like core-sheath microstructure formed at 70 wt% PEO (Figure 4.3e), the<br />

increased viscosity and correspondingly low mobility of PEO prevents an intact core-sheath<br />

structure from forming and phase separation is not sharply delineated, with a very large � ><br />

125


100 nm. At 80 wt% PEO (Figure 4.3f), well-defined domains of P2VP with � ≈ 26 nm are<br />

visible within the PEO matrix. Even at this low concentration of hydrophobic P2VP, the<br />

majority of the SPIONs prefer to reside within this phase.<br />

In summary, the SPIONs prefer to reside in the hydrophobic P2VP phase. Since they are<br />

coated with oleic acid, relatively little aggregation is observed in the microfibers.<br />

Comparison of Figures 4.3 and 4.4 suggests that the addition of SPIONs improves the extent<br />

of phase separation between PEO and P2VP without affecting the morphology that develops<br />

at a given blend composition. In this regard, the SPIONs act as a decompatabilizer between<br />

the two homopolymers. Such behavior has been previously reported 4 due to incompatability<br />

between the surface ligand and the host polymer phase. Some of the morphologies seen in the<br />

microfibers are attributed to differences in viscosity. To ascertain the effect of SPIONs on the<br />

viscosity of PEO and P2VP in chloroform, we have measured the zero-shear viscosity (�) of<br />

the following systems in chloroform at ambient temperature: 8 wt% P2VP and 2 wt% PEO,<br />

both with and without 1 wt% SPION. Without nanoparticles, the PEO solution possesses a<br />

higher viscosity (0.420 Pa-s) than the P2VP solution (0.201 Pa-s) despite the 4x higher<br />

polymer concentration in the latter. Incorporation of the SPIONs consistently lowers � in<br />

both systems and causes their viscosities to become comparable: 0.083 Pa-s for PEO and<br />

0.106 Pa-s for P2VP. A nanoparticle-induced reduction in viscosity has been previously<br />

reported for polymer melts, 43,44 and molecular dynamics simulations 45 have confirmed that a<br />

viscosity decrease upon nanoparticle addition is characteristic of repulsive systems. The<br />

solubility parameter of the oleic acid at 15.6 MPa1/2 is sufficiently different from that<br />

126


corresponding to the host P2VP, which is consistent with a repulsive system and an<br />

accompanying reduction in �. This observation also implies that the high-molecular-weight<br />

P2VP chains responsible for � tend to physisorb on the nanoparticle surface and contribute<br />

much less to �. 43<br />

To elucidate the chain packing within the morphologies observed in the electrospun<br />

microfibers of PEO/P2VP blends, we have used XRD to examine the crystalline structure of<br />

microfiber mats with and without added SPIONs. The reference XRD pattern of PEO powder<br />

clearly displays the two signature peaks of the monoclinic lattice of crystalline PEO at 19.06°<br />

2� (Peak I) and 23.22° 2� (Peak II). 46 Peak I corresponds to the (120) reflection, while peak<br />

II refers to the (112) and (032) reflections. 47 The average crystallite size (�) can be calculated<br />

by the Scherrer equation, given by 48<br />

� � K�<br />

�cos�<br />

where K is the shape factor (= 0.9 in the present case), β is a measure of line broadening (i.e.,<br />

full-width at half the ��maximum<br />

intensity) and � denotes the Bragg angle. The corresponding<br />

d-spacing is determined by the Bragg equation: 49<br />

d � n�<br />

2sin�<br />

in which n is the integral diffraction order equal to unity here. Pure PEO consists of large,<br />

well-defined crystals<br />

��<br />

(with � = 28.27 nm from Peak I) wherein the intensity of Peak I is<br />

~20% higher in intensity than Peak II. Values of � and the Peak I/Peak II ratio (I/II) extracted<br />

from the scattering profiles provided in Figure 4.4 for electrospun microfibers at different<br />

127<br />

(1)<br />

(2)


lend compositions are listed in Table 4.1. When PEO is electrospun, the positions remain<br />

nearly the same for both Peaks I and II, but peak broadening yields an increase in � and thus<br />

a reduction in �. This result has been previously reported 40 for electrospun systems since the<br />

timescale of crystal formation during electrospinning is much shorter 31 than that in a bulk<br />

system, such as the powder produced upon polymerization. A decrease in the I/II peak ratio<br />

to 0.93 means that the (120) reflection is stronger, which occurs when the PEO polymer<br />

chains preferentially orient along the axis of elongational flow during electrospinning. In this<br />

scenario, the PEO lamellae lie perpendicular to the fiber axis. 50<br />

As the P2VP content in the fiber increases (without SPIONs), several trends become<br />

apparent from the profiles displayed in Figure 4.4. The first is that the positions of Peaks I<br />

and II do not change within experimental error, which translates into a crystal d-spacing that<br />

is independent of blend composition. This observation is consistent with a PEO phase that<br />

remains relatively pure (unmixed) in the presence of P2VP at PEO concentrations as low as<br />

40 wt%. At lower concentrations, the peaks shift and become distorted, indicating enhanced<br />

phase mixing as inferred earlier. The values of t extracted from Peak I systematically<br />

decrease from ~28 to ~19 nm with increasing P2VP loading, indicating that the length of the<br />

crystals along the fiber axis decrease as the reservoir of PEO within the microfiber is lowered<br />

and the domains of PEO within the microfibers become smaller and more dispersed. In<br />

contrast, the width of the crystals oriented normal to the fiber axis fluctuates modestly<br />

between about 9 and 13 nm. Values of the I/II peak ratio are always less than unity in<br />

electrospun microfibers varying in blend composition, in which case the crystalline lamellae<br />

128


prefer lying perpendicular to the fiber axis in electrospun microfibers. Other than the diffuse<br />

interfaces discussed earlier, there is no evidence from XRD that the PEO and P2VP interact<br />

to any discernible extent. 51<br />

Addition of SPIONs measuring 18 nm in diameter at a concentration of a 2.5 vol% to the<br />

electrospun PEO/P2VP microfibers yields the XRD patterns presented in Figure 4.4.<br />

Pertinent metrics extracted from these profiles are provided in Table 4.2. In microfibers<br />

composed of 100 wt% PEO, the SPIONs promote a significant reduction (by about 30%) in<br />

crystal size along the fiber axis and almost no change in either crystal d-spacing or crystal<br />

size normal to the fiber axis. In addition, the I/II peak ratio is less than unity (0.46), which<br />

confirms that the PEO chains, as well as their innate ability to crystallize, are strongly<br />

affected by the presence of SPIONs within PEO. This dependence is not evident from the<br />

TEM images in Figure 4.3 since the images only show PEO/P2VP blend morphologies with<br />

SPIONs. Upon reducing the PEO concentration to 80 wt%, the longitudinal crystal size<br />

increases, but nonetheless remains lower than that in the absence of SPIONs. Further<br />

decreasing the PEO concentration results in a systematic reduction in longitudinal crystal<br />

size, as was seen in PEO/P2VP microfibers without SPIONs. These results, as well as the<br />

corresponding transverse crystal sizes, are shown as a function of blend composition in<br />

Figure 4.6 and confirm that, even in the PEO/P2VP blends, the SPIONs affect the crystal<br />

characteristics of PEO. Recall from TEM images such as those in Figure 4.3, however, that<br />

the SPIONs are preferentially sequestered within the P2VP phase.<br />

129


The influence of SPIONs on PEO in electrospun microfibers composed of PEO/P2VP<br />

blends can be explained in one of two ways. Firstly, a fraction of added SPIONs, while<br />

thermodynamically attracted to the P2VP phase due to the presence of the oleic acid ligands,<br />

remains kinetically trapped in the PEO phase during microfiber formation. Alternatively,<br />

some SPIONs reside along the PEO/P2VP interface. In both cases, the SPIONs contact PEO<br />

chains and affect their ability to enter into three-dimensional registry and form crystals. To<br />

decide which situation is more likely, we turn our attention to the I/II peak ratio, which is<br />

always less than unity in PEO/P2VP systems (except for pure PEO) without SPIONs.<br />

Confinement due to P2VP at any loading level appears to favor the formation of PEO<br />

lamellae oriented perpendicular to the fiber direction (transverse). With SPIONs, even pure<br />

electrospun PEO exhibits a peak ratio that is less than unity. As the concentration of PEO is<br />

reduced in this series, however, the peak ratio increases beyond unity so that the PEO<br />

lamellae are oriented parallel to the fiber direction (longitudinal). This change in lamellar<br />

orientation implies that the PEO phase is nearly pure (since the presence of SPIONs in PEO<br />

promote the opposite orientation), in which case the SPIONs promote decompatibilization, as<br />

concluded earlier. Since the PEO chains adopt a crystal orientation that is reminiscent of pure<br />

PEO but remain affected by SPIONs preferentially located within the P2VP phase, it<br />

immediately follows that SPIONs located along the PEO/P2VP interface are most likely<br />

responsible for the observed changes in the PEO crystal structure. This scenario is<br />

schematically illustrated in Figure 4.7. At 20 wt% PEO, the XRD pattern in Figure 4.5 shows<br />

a mostly amorphous scattering signature, which reflects the low loading level of semi-<br />

130


crystalline PEO. A broad Peak I and virtually nonexistent Peak II reveal very small crystals,<br />

which is consistent with the trend observed in Figure 4.6. In this case, phase mixing between<br />

PEO and P2VP and the existence of SPIONs within PEO hinder the crystallization of PEO<br />

chains.<br />

Conclusions<br />

Blends composed of a hydrophobic (P2VP) and a hydrophilic (PEO) polymer have been<br />

modified with SPIONs and electrospun into microfibers to discern the feasibility of<br />

controlling the spatial location of SPIONs on the basis of thermodynamic compatibility and<br />

blend morphology. At blend compositions favoring phase separation into dispersed domain<br />

or core-sheath morphologies, the SPIONs tend to locate within the P2VP phase. Analysis of<br />

the PEO crystallinity demonstrates that the addition of P2VP to electrospun microfibers<br />

without SPIONs generally reduces the size of PEO crystals and changes the orientation of<br />

PEO lamellae from longitudinal to transverse due most likely to confinement effects. In the<br />

presence of SPIONs, a similar, but more pronounced reduction in crystal size is observed<br />

with increasing P2VP concentration. In addition, the orientation of PEO lamellae changes<br />

from transverse to longitudinal, suggesting that SPIONs located along the PEO/P2VP<br />

interface influence PEO chain packing and crystallization.<br />

Acknowledgments<br />

This work was supported by a the NC Space Grant and the National Science Foundation<br />

Graduate Research Fellowship (K. E. R.), a National Aeronautics and Space Administration<br />

131


Graduate Fellowship (K. E. R.), and the NATO Science for Peace Program (L. M. B) and<br />

Indiana University Faculty Research Program (L. M. B.).<br />

132


Tables<br />

Table 4.1: XRD characteristics of PEO powder and electrospun PEO/P2VP microfibers.<br />

System Peak Angle<br />

(° 2�)<br />

PEO Powder<br />

100 wt% PEO<br />

Fiber Mat<br />

80 wt% PEO<br />

Fiber Mat<br />

60 wt% PEO<br />

Fiber Mat<br />

40 wt% PEO<br />

Fiber Mat<br />

133<br />

d-spacing<br />

(nm)<br />

��<br />

(nm)<br />

I 19.06 0.47 28.27<br />

II 23.22 0.38 12.27<br />

I 19.06 0.47 28.23<br />

II 23.25 0.38 11.22<br />

I 19.08 0.47 26.56<br />

II 23.22 0.38 10.17<br />

I 19.21 0.46 23.24<br />

II 23.40 0.38 12.78<br />

I 19.07 0.47 19.20<br />

II 23.34 0.38 9.40


Table 4.2: XRD characteristics of electrospun PEO/P2VP microfibers with SPIONs<br />

measuring 18 nm in diameter and added at a concentration of 2.5 vol%.<br />

System Peak Angle<br />

(°2�)<br />

100 wt% PEO<br />

Fiber Mat<br />

80 wt% PEO<br />

Fiber Mat<br />

60 wt% PEO<br />

Fiber Mat<br />

40 wt% PEO<br />

Fiber Mat<br />

20 wt% PEO<br />

Fiber Mat<br />

134<br />

d-spacing<br />

(nm)<br />

��<br />

(°)<br />

��<br />

(nm)<br />

I 19.27 0.46 0.28 19.62<br />

II 23.52 0.38 0.74 11.56<br />

I 19.13 0.46 0.35 23.56<br />

II 23.31 0.38 0.76 11.27<br />

I 19.16 0.46 0.40 20.79<br />

II 23.35 0.38 0.72 11.85<br />

I 19.07 0.47 0.59 14.09<br />

II 23.36 0.38 1.19 7.19<br />

I 19.61 0.45 7.30 1.14<br />

II — — — —


Figures<br />

Figure 4.1. TEM image of SPIONs measuring 16.4 nm in diameter and drop cast from<br />

chloroform.<br />

135


a.<br />

b.<br />

Figure 4.2. (a) SEM image of SPION-containing PEO/P2VP microfibers composed of 80<br />

wt% PEO and electrospun from an 8.5 wt% solution in chloroform. (b) An enlargement<br />

showing the surface of the microfibers included in (a). The inset in (b) displays a SPION-rich<br />

bead, the scalemarker corresponds to 500 nm.<br />

136


Figure 4.3. TEM images of SPIONs, measuring 18 nm in diameter at a concentration of 2.5<br />

vol%, in biphasic microfibers composed of PEO/P2VP at different PEO concentrations<br />

(labeled). Inset scalebars all represent 200 nm and all other scalebars represent 500 nm.<br />

137


16 18 20 22 24 26 28 30<br />

2��(degrees)<br />

138<br />

80% PEO<br />

100% PEO<br />

PEO Powder<br />

60% PEO<br />

Intensity (a.u.)<br />

40% PEO<br />

Figure 4.4. XRD patterns acquired from PEO powder and electrospun microfibers composed<br />

of PEO/P2VP at different PEO concentrations (labeled).


16 18 20 22 24 26 28 30<br />

2��(degrees)<br />

139<br />

80% PEO/NP<br />

80% PEO/NP<br />

40% PEO/NP<br />

60% PEO/NP<br />

Intensity (a.u.)<br />

20% PEO/NP<br />

Figure 4.5. XRD patterns acquired from electrospun microfibers composed of PEO/P2VP<br />

with SPIONs (18 nm and 2.5 vol%) at different PEO concentrations (labeled).


� (nm)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

100<br />

Longitudinal<br />

Transverse<br />

80<br />

Figure 4.6. Average PEO crystal size (t) extracted from XRD patterns and presented as a<br />

function of PEO concentration parallel (circles) and perpendicular (triangles) to the fiber axis<br />

for systems without (open symbols) and with (filled symbols) SPIONs. Values measured<br />

from PEO powder are included (triangles). The solid and dashed lines serve as guides for the<br />

eye.<br />

140<br />

60<br />

40<br />

PEO concentration (wt%)<br />

20


a.<br />

b.<br />

Figure 4.7. Schematic illustration depicting the arrangement of polymer chains in a<br />

core/sheath microstructure of PEO/P2VP (a) before and (b) after SPION addition (not to<br />

scale). Addition of SPIONs promotes a reduction in crystal size but a more parallel chain<br />

arrangement with respect to the fiber axis.<br />

141


References<br />

(1) Hamley, I. W. Angewandte Chemie-International Edition 2003, 42, 1692.<br />

(2) Maity, S.; Downen, L. N.; Bochinski, J. R.et al. Polymer 2011, 52, 1674.<br />

(3) Rabani, E.; Reichman, D. R.; Geissler, P. L.et al. Nature 2003, 426, 271.<br />

(4) Balazs, A. C.; Emrick, T.; Russell, T. P. Science 2006, 314, 1107.<br />

(5) Bates, F. S. Science 1991, 251, 898.<br />

(6) Bates, F. S.; Fredrickson, G. H. Physics Today 1999, 52, 32.<br />

(7) Mackay, M. E.; Tuteja, A.; Duxbury, P. M.et al. Science 2006, 311, 1740.<br />

(8) Rittigstein, P.; Torkelson, J. M. Journal of Polymer Science Part B-Polymer Physics 2006,<br />

44, 2935.<br />

(9) Reneker, D. H.; Chun, I. Nanotechnology 1996, 7, 216.<br />

(10) Doshi, J.; Reneker, D. H. Journal of Electrostatics 1995, 35, 151.<br />

(11) Taylor, G. Proceedings of the Royal Society of London Series A 1969, 313, 453.<br />

(12) Shin YM, H. M., Brenner MP Applied Physics Letters 2001, 78, 1149.<br />

(13) Jin, H. J.; Fridrikh, S. V.; Rutledge, G. C.et al. Biomacromolecules 2002, 3, 1233; Kenawy,<br />

E. R.; Bowlin, G. L.; Mansfield, K.et al. Journal of Controlled Release 2002, 81, 57.<br />

(14) Zhang, Q.; Wu, D. H.; Qi, S. L.et al. Materials Letters 2007, 61, 4027.<br />

(15) Sun, Z. C.; Zussman, E.; Yarin, A. L.et al. Advanced Materials 2003, 15, 1929.<br />

(16) Kalra, V.; Lee, J. H.; Park, J. H.et al. Small 2009, 5, 2323.<br />

(17) Muller, K.; Quinn, J. F.; Johnston, A. P. R.et al. Chemistry of Materials 2006, 18, 2397.<br />

(18) Bognitzki, M.; Frese, T.; Steinhart, M.et al. Polymer Engineering and Science 2001, 41, 982.<br />

(19) Li, D.; Ouyang, G.; McCann, J. T.et al. Nano Letters 2005, 5, 913.<br />

142


(20) Jin, H. J.; Fridrikh, S.; Rutledge, G. C.et al. Abstracts of Papers of the American Chemical<br />

Society 2002, 224, U431.<br />

(21) Sun, X. Y.; Shankar, R.; Borner, H. G.et al. Advanced Materials 2007, 19, 87.<br />

(22) Mitchell, S. B.; Sanders, J. E. Journal of Biomedical Materials Research Part A 2006, 78A,<br />

110.<br />

(23) Wei, M.; Lee, J.; Kang, B. W.et al. Macromolecular Rapid Communications 2005, 26, 1127.<br />

(24) Li, D.; Xia, Y. N. Nano Letters 2004, 4, 933.<br />

(25) Jiang, H. L.; Hu, Y. Q.; Li, Y.et al. Journal of Controlled Release 2005, 108, 237.<br />

(26) Greiner, A.; Wendorff, J. H.; Yarin, A. L.et al. Applied Microbiology and Biotechnology<br />

2006, 71, 387.<br />

(27) Valiquette, D.; Pellerin, C. Macromolecules 2011, 44, 2838.<br />

(28) Chen, M. L.; Dong, M. D.; Havelund, R.et al. Chemistry of Materials 2010, 22, 4214.<br />

(29) Zhang, Y. Z.; Huang, Z. M.; Xu, X. J.et al. Chemistry of Materials 2004, 16, 3406.<br />

(30) Kim, G. M.; Wutzler, A.; Radusch, H. J.et al. Chemistry of Materials 2005, 17, 4949.<br />

(31) Wang, A.; Singh, H.; Hatton, T. A.et al. Polymer 2004, 45, 5505.<br />

(32) Wang, B. B.; Li, B.; Xiong, J.et al. Macromolecules 2008, 41, 9516.<br />

(33) Song, T.; Zhang, Y. Z.; Zhou, T. J.et al. Chem. Phys. Lett. 2005, 415, 317.<br />

(34) Bronstein, L. M.; Huang, X.; Retrum, J.et al. Chem. Mater. 2007, 19, 3624.<br />

(35) Bronstein, L. M.; Atkinson, J. E.; Malyutin, A. G.et al. Langmuir 2011, 27, 3044.<br />

(36) Shtykova, E. V.; Huang, X.; Remmes, N.et al. J. Phys. Chem. B 2007, 111, 18078.<br />

(37) Polymer Data Handbook; Oxford University Press: New York, New York, 1999.<br />

(38) Park, S. C.; Kim, B. J.; Hawker, C. J.et al. Macromolecules 2007, 40, 8119.<br />

(39) Yahya, M. Z. A.; Arof, A. K. European Polymer Journal 2003, 39, 897.<br />

143


(40) Huang, J. X.; Virji, S.; Weiller, B. H.et al. Journal of the American Chemical Society 2003,<br />

125, 314.<br />

(41) Reneker, D. H.; Yarin, A. L.; Fong, H.et al. Journal of Applied Physics 2000, 87, 4531.<br />

(42) D.J. Lohse, T. P. R., L.H. Sperling Interfacial Aspects of Multicomponent Polymer Materials;<br />

John Wiley & Sons: New York, NY, 1995.<br />

(43) Jain, S.; Goossens, J. G. P.; Peters, G. W. M.et al. Soft Matter 2008, 4, 1848.<br />

(44) Mackay, M. E.; Dao, T. T.; Tuteja, A.et al. Nature Materials 2003, 2, 762.<br />

(45) Smith, G. D.; Bedrov, D.; Li, L. W.et al. Journal of Chemical Physics 2002, 117, 9478.<br />

(46) Tadokoro, H.; Chatani, Y.; Yoshihara, T.et al. Makromolekulare Chemie 1964, 73, 109.<br />

(47) Bortel, E.; Hodorowicz, S.; Lamot, R. Makromolekulare Chemie-Macromolecular Chemistry<br />

and Physics 1979, 180, 2491.<br />

(48) P, S. Gottinger Nachrichten Gesell. 1918, 2, 98.<br />

(49) Bragg, W. L. Philosophical Magazine 1920, 40, 169.<br />

(50) Zhang, J.-F.; Yang, D.-Z.; Xu, F.et al. Macromolecules 2009, 42, 5278.<br />

(51) Mohapatra, S. R.; Thakur, A. K.; Choudhary, R. N. P. Journal of Power Sources 2009, 191,<br />

601.<br />

144


CHAPTER V<br />

Nanostructured Organometallic Polymer Systems Containing<br />

Poly(ferrocenylsilanes)<br />

5.1 Introduction<br />

An organometallic polymer contains a transition metal in the main chain or more<br />

specifically has a metal-carbon σ or π bond. 1 The term ‘transition metal’ refers to any<br />

element that has an incomplete d sub-shell or which can give rise to cations with an<br />

incomplete d sub-shell. 2 The properties that make these transition metals interesting,<br />

however, are their ability to have multiple oxidation states that can be easily controlled<br />

through electric currents. The first organometallic polymer was synthesized in 1955 as<br />

poly(vinylferrocene) and displayed reversible oxidation-reduction properties. 3 The<br />

incorporation of metallic elements into polymer systems allow different coordination<br />

numbers and geometries and thus supply valuable magnetic, optical or catalytic properties.<br />

Structurally, metallopolymers can be divided into three categories: metals incorporated<br />

directly into the polymer chain, π or σ-coordinated metals, and metallic moieties pendant to<br />

the polymer backbone or in side chains. 4 Some challenges that have thwarted the synthesis of<br />

organometallics include low molecular weights, oligomers, impurities, and insolubility.<br />

In the early 1990s, Manners et al. 5 fabricated poly(ferrocenylsilane) (PFS) with a high<br />

molecular weight and narrow polydispersity through a ring-opening polymerization method,<br />

145


and this synthesis technique has since yielded other strained monomers including bridging<br />

elements such as germanium, tin and phosphous. 6 Poly(ferrocenylsilane)s can be tuned to<br />

adopt either a semi-crystalline or amorphous state as a consequence of the constituent groups<br />

on the silicon atom 7 where symmetrically substituted constituents, R=R'=Me, impart<br />

crystallinity. The prevalence of iron in the main chain conveys some interesting properties<br />

not present in non-metal containing polymers such as redox-activity due to the reversible<br />

(Fe(II)/Fe(III) couple) 2 , the ability to be pyrolyzed into a magnetic ceramic, 8 and semi- and<br />

photo-conductivity. 9<br />

Block copolymers are macromolecules containing two or more types of repeat units<br />

within long contiguous sequences, or “blocks,” of the same unit. These sequences are<br />

covalently linked to form a single macromolecule. They can spontaneously self-organize, or<br />

microphase-separate, into a variety of ordered nanoscale morphologies: lamellae,<br />

hexagonally packed cylinders, spherical micelles on a body- or face-centered cubic lattice, or<br />

complex bicontinuous morphologies (e.g., the gyroid). 10 The type of morphology exhibited<br />

by the block copolymer depends on the chemical attributes and lengths of the blocks, the<br />

molecular architecture and temperature, as well as the presence of an additive such as a<br />

solvent, 11 homopolymer, 12 nanoparticle, 13 or another block copolymer. 14 Due to their ability<br />

to microphase-separate, block copolymers constitute a versatile platform for a number of<br />

existing and emerging technologies such as adhesives, membranes, drug delivery,<br />

biomaterials and nanolithography. 15<br />

146


The formation of metal-containing block copolymers was possible due to the<br />

controllability of anionic ring-opening polymerization of silicon-bridged ferrocenes. 16<br />

Polymerization must occur through sequential addition of monomers with decreasing end-<br />

group reactivity such that poly(styrene) (PS) ~ poly(isoprene) (PI) > PFS ><br />

poly(dimethylsiloxane) (PDMS). 16 The first two PFS-containing block copolymers, 17 PS-b-<br />

PFS and PFS-b-PDMS, have allowed the synthesis of block copolymers of PFS and PI, 18<br />

poly(methylmetacrylate) (PMMA), 19 poly(2-vinyl pyridine) (P2VP),<br />

poly(ferrocenylethylmethylsilane) (PFEMS), 20 and hybridization with polypeptides. 21 These<br />

metallated block copolymers have been shown to self-assemble in the solid state and have<br />

ordered into spherical, cylindrical, lamellar, and gyroid morphologies. 22 Triblock copolymers<br />

can be selectively etched so that only PFS is remaining, as seen in Figure 5.3 which gives<br />

rise to many lithographic applications. In the solid state, amorphous PFS with asymmetrical<br />

constituents on the silicon atom can be used to promote phase separation.<br />

Periodic PFS domains can be converted to iron-rich clusters within a ceramic domain by<br />

pyrolysis at temperatures of about 600°C, which can lead to the growth of single-walled<br />

carbon nanotubes for soft lithography, 23 the formation of magnetic ceramics 24 and PFS-<br />

derived catalysts. 25 Solution self-assembly of diblock copolymers occurs when a block-<br />

selective solvent is utilized to induce segregation of the solvent-incompatible ‘core’<br />

surrounded by the solvent-compatible ‘corona.’ Micelles of PFS-containing block<br />

copolymers have produced micellar morphologies such as cylinders, 26,27 tubes, 27 fibers, 28 and<br />

tapes. 18 Investigations performed in the bulk typically utilize poly(ferrocenyldimethylsilane)<br />

147


(PFDMS), whose crystalline structure is responsible for the growth of micelles. 29 Initial<br />

experiments of PFDMS-b-PDMS (with a block ratio of 6:1, respectively) in hexanes<br />

demonstrated that cylinders with a PFS core and a PDMS corona formed. 26 Imaging these<br />

organometallic, self-assembled structures by transmission electron microscopy (TEM) is<br />

facilitated due to the inherent contrast arising from the iron-rich core. Forming a core of<br />

PFDMS in cylindrical micelles requires a corona:core ratio of at least 5:1, 18,26 but when this<br />

ratio reaches 12:1, hollow, tube-like structures most likely form. 19,27 The length of the<br />

cylindrical micelles can also be controlled: the addition of a small amount of a common<br />

solvent will allow growth of the micelles, whereas ultrasound or high temperatures will<br />

cleave long cylindrical micelles and yield short ones instead. Self-assembly of PFDMS-b-<br />

P2VP has been investigated in several alcohols, and cylindrical micelles are observed to form<br />

in isopropanol but not ethanol. 30 The reason for this is attributed to the fact that isopropanol<br />

is a better solvent for PFDMS than ethanol on the basis of solubility parameters. Thus, the<br />

PFS chains remain solvated longer and have time to rearrange and crystallize more<br />

efficiently. In addition, PFS-b-PI block copolymers self-organize into cylindrical micelles in<br />

a PI-selective solvent, and the vinyl groups in the PI corona were subjected to a Pt(0)-<br />

catalyzed crosslinking reaction, which was a precursor to making PFS nanoceramics with<br />

shape retention by pyrolysis. 31 Nanocomposite self-assembled structures consisting of PFS-b-<br />

PVMS wormlike micelles react with Ag[PF6] to create a one-dimensional array of silver<br />

nanoparticles encapsulated within the worm-like micelles, 32 thereby allowing for alignment<br />

of highly oriented nanoparticle arrays.<br />

148


5.2 Experimental<br />

5.2.1 Materials<br />

Characteristics of all the polymers employed here, including their chemical constitution,<br />

number-average molecular weight (Mn), polydispersity index (PDI), and composition (�PFS),<br />

are listed in Table 5.1. Reagent-grade chloroform and dimethylformamide (DMF) were<br />

obtained from Fisher Scientific (Fairlawn, NJ). Tetrahydrofuran (THF) was purchased from<br />

Fisher Scientific (Fairlawn, NJ), and dichloromethane (DCM) was acquired from Sigma-<br />

Aldrich (Milwaukee, WI)<br />

5.2.2. Synthesis of Specialty Polymers<br />

All organometallic polymers were synthesized in the Manners laboratory at Bristol<br />

University specifically for this research. Due to this, molecular weights and PDI’s are not<br />

consistent between samples, and extremely limited amounts of polymer prohibit<br />

determination of solution properties. A month-long visit to Bristol University was undertaken<br />

in 2008 to prepare low-molecular-weight poly(ferrocenylpropylmethylsilane) (PFPMS) and<br />

poly(ferrocenyl-dimethylsilane)-b-polyisoprene(PFS-b-PI).<br />

Dilithioferrocene·tetramethylethylenediamine was previously prepared in the Manners<br />

laboratory via the synthesis by Rider. 20 All parts of the synthesis were conducted under<br />

nitrogen since ferrocene can self-combust. Dilithioferrocene·tetramethylethylenediamine<br />

(15.4 g) was suspended in diethyl ether (500 mL) and the suspension was chilled to -78°C in<br />

an acetone/dry ice bath. Distilled dichloroethylmethylsilane (8.4g) was added dropwise via<br />

149


canulla tubing to the flask, which was allowed to warm to ambient temperature for 4 h.<br />

During this heating, the color of the material changed from orange to dark red. The solvent<br />

was then evaporated and the solids were dried under dynamic vacuum overnight. The solids<br />

were then dissolved in hexanes and recrystallized twice at -55°C to remove any impurities.<br />

This was then followed with two sublimation cycles to eliminate any reacted ferrocenophane.<br />

In an inert glove box, 250 mg of the ethylmethyl-silaferrocenophane was dissolved in 5 mL<br />

of THF to which 1.5 μL of n-butyllithium was added (1.6 M solution in hexanes). The<br />

polymerization was allowed to proceed for 30 min and was terminated by degassed<br />

methanol. The solid was then precipitated into methanol and dried overnight under dynamic<br />

vacuum, resulting in PFEMS, as depicted in Figure 5.1. Thermal ring-opening<br />

polymerization, as described above, is diverse in that it allows a range of silicon-substituted<br />

functional groups. Thus, PFDMS and poly(ferrocenylphenylmethylsilane) (PFPMS) can be<br />

prepared by the same synthetic scheme with different starting monomers.<br />

5.2.3. Preparation of Nanoparticles<br />

The thermal decomposition of iron oleate with oleic acid occurred through thermal<br />

decomposition and resulted in iron oxide nanoparticles with diameters ranging from 10-14<br />

nm. 33,34 These as-synthesized nanoparticles are crystalline according to x-ray diffraction<br />

(XRD) and contain mostly wüstite (Fe(1-x)O) and some spinel (most likely Fe3O4), 33 and are<br />

superparamagnetic iron oxide nanoparticles (SPIONs) according to magnetic<br />

measurements. 35 The resultant SPIONs were precipitated with a solvent mixture and then<br />

150


separated by centrifugation and suspended in chloroform at concentrations between 10 and<br />

50 mg/mL<br />

5.2.4. Preparation of Electrospun Fibers<br />

Suspensions for electrospinning were prepared by first dissolving PFS in either 9:1<br />

THF:DMF or DCM and then adding between 0.5 – 4% Triton-X 405 surfactant. The in-house<br />

electrospinning unit, operated at 8 kV, employed a syringe pump and an Al collection target.<br />

The separation distance was held constant at 12 cm, and the solution flow rate varied from 10<br />

to 35 �L/min. Electrospun fibers were collected on the Al plate, as well as on carbon-coated<br />

transmission electron microscopy (TEM) grids adhered to the plate.<br />

5.2.5. Characterization of PFS Nanomaterials<br />

Transmission electron microscopy was performed on a field-emission Hitachi HF2000<br />

microscope operated at 200 kV. Complementary scanning electron microscopy (SEM)<br />

images were collected from specimens sputter-coated with 6 nm of Au/Pd on a JEOL 6400F<br />

field-emission microscope operated at 5 kV. X-ray diffractometry (XRD) was performed on<br />

a Rigaku Smartlab diffractometer with CuKα radiation at 2� angles ranging from 5 to 30° in<br />

0.01° increments at a wavelength (�) of 0.1541 nm.<br />

5.3 Electrospinning and Characterization of PFS Homopolymers<br />

Functional materials can be created through polymer nanocomposites in order to combine<br />

different sets up properties. One way these materials can be formed is through<br />

electrospinning, the application of electrostatic forces to a polymer solution to create<br />

151


nano/micro-fibers with characteristics that can be tailored by processing, solution, and<br />

polymer properties. 36 Functionality must be added to the polymer however; often through<br />

the use of dispersed metallic nanoparticles, 37 surface modification, 37 or post-spinning<br />

annealing techniques. 38<br />

The shortcomings of all of these modification techniques, however, are pre- and/or post-<br />

processing steps, as well as aggregation issues associated with metallic materials. A<br />

straightforward solution to these disadvantages is to combine the favorable properties of<br />

metallic and polymer materials into one molecular species: an organometallic polymer. The<br />

goal of this section is to lay the groundwork for electrospinning organometallic polymers and<br />

investigate how the processing conditions during electrospinning influence the resulting fiber<br />

morphology.<br />

As previously mentioned, PFS can be synthesized to exist in either a semi-crystalline or<br />

amorphous state due to the constituent groups on the silicon atom. 7 Symmetrically<br />

substituted constituents, R=R'=Me, tend to impart crystallinity. The electrospinning of<br />

polyferrocenylalkylsilanes was first reported by Chen et al. 39 in the case of PFDMS with a<br />

molecular weight of 95.6 kDa in a 9:1 volume ratio of THF:DMF. Initially, replication of<br />

their solution/processing conditions with PFDMS at 95.6 kDa did not result in similarly<br />

smooth fibers. Concentrations of 15-20% PFDMS in a 9:1 volume ratio of THF:DMF yielded<br />

fibers with a large population of elongated beads along with a wide distribution of fiber<br />

diameters. At concentrations higher than 15 wt% DMF, which was increased to raise the<br />

conductivity of the solution, the PFDMS precipitated out of solution. The addition of<br />

152


surfactants have been shown to lower the bead density in electrospun fibers by reducing<br />

surface tension and increasing conductivity. 40 The surfactant, Triton X-405 (polyethylene<br />

glycol tert-octylphenyl ether), was added at concentrations ranging from 0.5 – 4 wt%. At<br />

concentrations higher than 2%, however, the surfactant segregated into domains within the<br />

polymer solution. For PFS systems, the addition of surfactant successfully eliminated the<br />

majority of bead defects and also resulted in a higher mean fiber diameter, as seen in Table<br />

5.2. Electrospinning PFDMS in DCM was also successful, which resulted in a more uniform<br />

distribution of fiber diameters, as well as residual bead. The PFPMS was electrospun from a<br />

9:1 THF:DMF solution with 2% Triton X-405 to delineate differences for the amorphous<br />

polymer. Representative SEM images of all the PFS fibers are displayed for comparison in<br />

Figure 5.2. Fibers were very heavily beaded, even with the addition of surfactant. Widely<br />

varying fiber diameters are responsible for the large standard deviations. A broad range of<br />

fiber diameters result from ‘splaying,’ wherein the axial forces at the tip of the spinneret are<br />

greater than the longitudinal force. For both crystalline and amorphous PFS, a narrow<br />

concentration window exists for spinnability. This may be due, in part, to the low molecular<br />

weights of the polymers employed, ~96 kDa. If the concentration is too low, the electrostatic<br />

forces dominate, and the jet breaks up into droplets or electrospraying ensues. On the other<br />

hand if the concentration is too high, viscous forces dominate, and the polymer solution<br />

begins to solidify at the tip of the syringe with little to no spraying. Another point of interest<br />

is that relatively high concentrations (~20 wt%) are required to produce fibers. This tells us<br />

that below a critical threshold (~18 wt% in the case of PFDMS), the polymer chains produce<br />

153


a relatively dilute solution. As the concentration increases and the distance between chains<br />

decreases, strong interactions between ferrocene groups lead to an abrupt increase in<br />

viscosity to the point where these interactions dominate over electrostatic forces.<br />

To elucidate the arrangement of chains comprising the observed morphologies, XRD has<br />

been used to explore the crystalline structure of the powders and fibers. As a reference point,<br />

XRD analysis of PFDMS powder reveals a strong principal peak at 13.95° 2�, which<br />

corresponds to an interplanar distance (d-spacing) of 0.63 nm. This value is comparable to<br />

that reported 42,43 previously for melt- and solid-state polymerized PFDMS and is indicative<br />

of the distance between adjacent planes of iron atoms, since this atom possesses a large<br />

scattering length. The average crystallite size (�) was obtained from the Scherrer equation,<br />

which can be expressed as 41<br />

� � K�<br />

�cos�<br />

and found to be 5.70 nm. Here, K represents the shape factor (= 0.9), β is the full-width at<br />

half the maximum intensity, �� and � is the Bragg angle. In addition, two smaller peaks located<br />

at 20.09° and 23.39° 2� in Figure 5.4 were also found to exist and are representative of<br />

tetragonal/hexagonal chain packing and indicative of a secondary crystalline phase in the<br />

PFDMS powder. When PFDMS is electrospun, the d-spacing associated with the primary<br />

peak increases slightly, most likely due to distortion of the chains as they undergo flow. The<br />

value of � corresponding to the principal peak, however, increases substantially from 5.74 to<br />

17.20 nm, which is contrary to what is normally observed in electrospun fibers. 42 As the<br />

154<br />

(1)


polymer chains are stretched and become elongated, packing between chains is marginally<br />

compromised and the ferrocene groups are pulled apart allowing for larger crystal sizes (see<br />

Figure 5.5).<br />

Organometallic nanocomposites are often fabricated by physically blending metal or<br />

metal oxide nanoparticles with polymer matrices. In this study, we investigate the effect of<br />

blending SPIONs with PFDMS to ascertain if any coordination takes place between the two<br />

sources of iron. A suspension of 10 nm SPIONs in chloroform was blended with PFDMS and<br />

subsequently electrospun. Information regarding the subsequent crystalline structure is<br />

provided in Table 4 and indicates a negligible increase in d-spacing by 0.01 nm and a<br />

decrease in �. While the SPIONs have little effect on chain packing, the reduction in t<br />

suggests that the SPIONs may be positioned between ferrocene groups so that the chains<br />

wrap around the nanoparticle, as depicted in the schematic diagram displayed in Figure 5.5.<br />

A consequence of this proposed explanation is that a larger SPION is expected to promote a<br />

decrease in d-spacing, since the local radius of curvature due to the SPION would be<br />

reduced. Increasing the SPION diameter to 14 nm does yield a decrease in d-spacing, which<br />

is lower than that of pure PFDMS fiber. Images collected from these systems confirm that<br />

little, if any, SPION aggregation occurs in these systems, which provides additional support<br />

that individual SPIONs are coordinated with the PFDMS chains.<br />

The effect of asymmetrical silicon atom substitution on fiber crystallinity in PFPMS has<br />

also been investigated in powder and fiber form. The powder displays an amorphous halo in<br />

XRD, which is not surprisingly on the basis of previous reports. 43 Of keen interest here are<br />

155


the crystalline peaks that become evident in XRD patterns collected from electrospun<br />

PFMPS fibers. These peaks evince that electrospinning can induce crystalline order in an<br />

otherwise amorphous polymer. Two broad peaks are visible at 12.23° and 16.71° 2�,<br />

corresponding to interplanar distances of 0.72 and 0.53 nm, respectively. Corresponding<br />

values of � are 20.10 and 20.50 nm, respectively. A d-spacing of 0.69 nm has been<br />

previously found 44 for the intrachain distance between ferrocene groups in PFEMS. The<br />

present system contains a phenyl ring rather than an ethyl group, in which case a larger d-<br />

spacing would be consistent. In addition, the interchain distance between ferrocene units is<br />

measured to be 0.52 nm, which agrees favorably with our value of 0.53 nm. These results<br />

demonstrate that shear-induced crystallinity can be imparted to otherwise amorphous PFS<br />

polymers and that chain alignment occurs within PFMPS electrospun fibers.<br />

5.4. Phase behavior of binary blends of PFS in Elastomeric Matrices<br />

Blending an AB diblock copolymer with homopolymer A (hA) can lead to improved<br />

emulsifying qualities 45 and interfacial elasticity 46 , since the copolymer effectively behaves as<br />

a macromolecular surfactant 47 . The blocks of such binary blends can form either dry or wet<br />

brushes depending on the value of MhA/MA (hereafter denoted α), where Mi denotes the<br />

molecular weight of species i (i = hA or A). When α is large, a ‘dry brush’ forms due to the<br />

entropic penalty that arises due to the inability of the homopolymer to penetrate the dense<br />

brush created by the copolymer block 48 . In extreme cases, this drives the system to<br />

macrophase-separate into discrete domains of homopolymer and copolymer. When α is of the<br />

order unity or smaller, a ‘wet brush’ develops wherein homopolymer molecules penetrate<br />

156


and swell the copolymer brush 49 . The morphologies of these miscible microphase-separated<br />

blends can be controllably altered by modifying 12 α, the blend composition, the copolymer<br />

composition, and the thermodynamic incompatibility (χN), which relates to temperature.<br />

Block-selective solvents serve as the low-molecular-weight limit of homopolymers and<br />

ensure that the compatible block is fully wetted and swollen. Within a block-selective<br />

solvent, diblock copolymer molecules aggregate so that the solvent-incompatible blocks are<br />

sequestered within a core surrounded by a swollen corona of the solvent-compatible block.<br />

Depending on the copolymer composition, the solvated corona typically possesses a larger<br />

volume than the collapsed core and is responsible for the curvature induced at the core-<br />

corona interface 31 .<br />

Although the formation of cylindrical micelles is not unusual, PFS micelles are unique in<br />

that their length is driven by crystallization, in which case it is possible to ‘grow’ the<br />

cylinders by thermal treatment. Such crystal-stabilized nanostructures can lead to<br />

temperature-responsive properties. If a specimen containing elongated crystalline micelles is<br />

heated and the melting point of PFS is surpassed, the micelles will melt and then reform<br />

classical morphologies. In the case of conductive cylinders composed of PFS block<br />

copolymers, heating would eliminate conductive pathways so that the cylinders could serve<br />

as a thermal sensor. In this section, we investigate the behavior of PFS-b-PI micelles in two<br />

different elastomeric matrices that are compatible with the PI corona and how cross-linking<br />

affects the copolymer morphology.<br />

157


5.4.1 PFDMS-b-PI/PI Blend by Solvent Casting<br />

According to Manners and co-workers, 31 PFS-b-PI micelles form cylinders in an<br />

isoprene-selective solvent, such as hexane, and they have reported that the lengths of these<br />

micelles range from 100 nm to 1 μm. The copolymer used here was synthesized by living<br />

anionic polymerization in THF at ambient temperature. The molecular weight and block ratio<br />

were determined using GPC and 1 H NMR. Hexane was used to assemble one-dimensional<br />

micellar structures, the morphology of which is governed by varying the PI:PFS block ratio.<br />

When the blocks of PI are much smaller than those of PFS, the block copolymers form tape-<br />

like lamellae. When the length of the PI block far exceeds that of PFS, however, cylinders<br />

form with a crystalline PFS core. 50 Furthermore, these micelles can be grown epitaxially<br />

through the addition of more polymer, which serves to emulate a living polymer growth<br />

reaction. 51 Cylindrical micelles develop by dissolving a PFDMS-b-PI block copolymer in<br />

hexane at elevated temperatures. In the present work, cylinders form when the sample is<br />

heated to 60°C for 30 min and then allowed to cool at ambient temperature. The lengths of<br />

the resultant cylindrical micelles depend on cooling time 51 and can exceed 1 �m with some<br />

end-to-end alignment, as seen in Figure 5.6. Note that the micelles, measuring about 14.9 nm<br />

across, appear to be clustered together with little separation between individual cylinders.<br />

5.4.2. Cross-linking within Polyisoprene<br />

Organic “nanowires” are useable in a wide range of plastic electronic materials such as<br />

solar cells and disposable circuits. 52 Traditional disadvantages of plastic electronics arise due<br />

to high costs and difficult processability. Micelles composed of PFS-b-PI block copolymers<br />

158


have many characteristics that make them favorable candidates for nanowires: a conductive<br />

core, insulting sheath, robust mechanical properties, nanoscale dimensions, and lengths<br />

exceeding 2 �m. The goal of the present work is to crosslink these ‘nanowire’ micelles into<br />

an elastomeric matrix to determine their feasibility for soft electronics. Polyisoprene is a<br />

natural choice for the matrix polymer due to the high degree of cross-linking that can be<br />

achieved because of the large number of unsaturated bonds available in the system. These<br />

bonds can be reacted with sulfur atoms to connect neighboring molecules via a 6-8 chain<br />

sulfur ‘bridge’. This fundamental reaction, known as vulcanization, 53 constitutes the standard<br />

chemical reaction to cross-link natural rubber and has been widely used in the rubber<br />

industry for over 100 years. One complication of this reaction is, however, the high<br />

temperature (~170°C) needed for this reaction to proceed. Since the melting point of PFS is<br />

128°C, the cylinders would evolve at temperatures required for vulcanization.<br />

For tests aimed at blending micelles with low concentrations of PI, 1 wt% PI was added<br />

to the micelle suspension in hexane and allowed to slowly dissolve quiescently. The polymer<br />

suspension was subsequently dropcast onto a TEM grid for direct imaging. In Figure 5.7a, a<br />

TEM image establishes that, at a ratio of 3:100 PFS-b-PI:PI, the dimensions of the cylinders<br />

remain unaffected by the addition of PI. While the cylinders form large aggregates, they are<br />

now separated by an average distance of ~57 nm, which is attributed to the interaction of the<br />

PI matrix with the corona of the cylinders. At the solution concentration employed, the<br />

resultant PI films is insufficiently thin to remain stable and consequently dewets to forms<br />

discrete islands. 54 When the ratio of PFS-b-PI:PI is increased to 3:25 (Figure 5.7b), similar<br />

159


ehavior is observed, and a dewetting pattern reminiscent of spinodal dewetting 55 is uniform<br />

across the grid. In this case, the cylinders are separated further, with an average intermicellar<br />

distance of 94 nm, and are less aggregated than at 3:100. Taken together, these results<br />

confirm that the addition of PI does not affect micelle morphology, but interactions between<br />

PI with the coronae of the micelles promotes greater separation and less aggregation of the<br />

micelles.<br />

Self-assembled PFDMS-b-PI cylinders that had been permitted to grow for 24 h have<br />

been added to a PI solution to disperse conductive organometallic cylinders into a cross-<br />

linked rubber matrix, as schematically depicted in Figure 5.8. The PI was cross-linked<br />

according to straightforward vulcanization with the materials listed in Table 5.4. The<br />

benchmark reaction was allowed to proceed under nitrogen at 110, 120 and 130°C without<br />

micelle addition for 3-4 h, until the surface was no longer tacky. Based on these results, 120<br />

°C was chosen as the reaction temperature since it was still less than the melting point of the<br />

PFS and yielded cross-linked PI. In the subsequent analysis, PI was added to the micelle<br />

suspension in hexane at a ratio of 1:1000 PFS-b-PI:PI and then placed in a vacuum oven.<br />

Due to the solvent that was not previously present, the cross-linking time was increased to 5<br />

h to allow solvent evaporation and subsequent cross-linking to occur. The resultant elastomer<br />

was cryomicrotomed to produce electron-transparent sections so that the internal structure of<br />

the micelles cross-linked in PI could be examined by TEM. The TEM images displayed in<br />

Figure 9 demonstrate that cylindrical micelles are no longer observed and what remains<br />

appear as iron particles ranging in size from 100-500 nm. Possible reasons for PFS-b-PI<br />

160


micelle disassociation are two-fold: (i) the temperature necessary to induce PI cross-linking<br />

was too close to the melting point of the micelles, or (ii) the micelles were not sufficiently<br />

stable to survive the long cross-linking time of the PI. Based on these observations, cross-<br />

linking must occur at significantly lower temperatures, which effectively eliminates PI as the<br />

choice of matrix polymer. In addition, shell-cross-linked PFS-b-PI cylinders will be used to<br />

produce a stable micelle capable of surviving cross-linking.<br />

5.4.3 Shell-Cross-Linking of Cylindrical PI-b-PFS Micelles<br />

To enhance the stability of the cylindrical micelles, the peripheral vinyl groups on the PI<br />

corona chains were cross-linked via a Pt(0)-catalyzed hydrosilylation reaction, which was<br />

conducted using a vinyldimethylsiloxy-terminated polydimethylsiloxane with triethoxysilane<br />

in the presence of Pt(0) catalyst in hexane at ambient temperature. Subsequent 1 H NMR<br />

analysis of the product indicated that all vinyl groups were consumed. The success of the<br />

hydrolsilylation cross-linking reaction was monitored by transferring micelles from a PI-<br />

selective solvent (hexane) to a neutral solvent (THF). 56 In the case of non-cross-linked<br />

samples, the aggregates disassociated in the neutral solvent. When an insufficient amount of<br />

cross-linker or catalyst was used and the shell-cross-linking reaction was incomplete, the<br />

micelles swelled substantially. Completion of the reaction resulted in no discernible swelling<br />

upon transfer to THF and no post-reaction growth. The only change in cylindrical<br />

morphology, as determined by TEM, is that the previously rigid cylinders appear to become<br />

more flexible. The shell-cross-linked micelles used here are 43 nm wide and range in length<br />

up to 1.5 μm, as evidenced by Figure 5.10.<br />

161


5.4.4. Cross-linked PVMS as the Matrix Polymer<br />

Poly(vinyl methyl siloxane) (PVMS) was chosen because it satisfied several conditions;<br />

solubility in hexane, an elastomer with a low Tg (~150 K), solubility parameter compatibility<br />

with the PI corona of the cylindrical micelles and the solvent hexane (Table 5.5), the ability<br />

to form a cross-linked network at temperatures below the Tg of the micelles, and the potential<br />

for interesting interactions between the vinyl group and the corona. PVMS was synthesized<br />

in house using an equilibrium-controlled ring opening polymerization (ROP) and step-growth<br />

polymerization methods. Initially, vinyl methyl dichlorosilane was dissolved in a dilute<br />

aqueous solution of HCl to form a mixture of cyclic silicone structures and short oligomeric<br />

vinyl methyl siloxane chains. In the second step, the cyclic structures were separated by<br />

vacuum distillation. The short-chain silanols were then condensed until the desired polymer<br />

chain was reached. This occurred under mild basic conditions resulting in high yields of<br />

PVMS, 95% and greater.<br />

Compatibility between the matrix polymer and the micelle corona is important in order<br />

to assist in good dispersion and maintaining their cylindrical shape. A dilute solution of 5<br />

wt% PVMS containing PFS-b-PI micelles was blended in order to investigate whether the<br />

cylindrical mcielles could hold their shape when dispersed within PVMS, the matrix polymer<br />

and dropcast on a carbon-coated TEM grid for TEM analysis. As evidenced by Figure 5.11,<br />

we see an uneven film of PVMS that is very thick in some areas and thin in others with<br />

thicker PVMS islands in these sections. We see that the majority of the diblock copolymer<br />

micelles exists within the thick PVMS sections, and even in the thinner areas that the<br />

162


micelles are surrounded by an additional ‘corona’ of PVMS. One would expect high<br />

interactions between PI and PVMS based on the fact that both contain reactive groups. This<br />

is evidence of excellent compatibility between the PVMS and the micelles. The most<br />

important result is the cylindrical shape is intact at lengths greater than 750 nm with little<br />

aggregation. In order to cross-link PVMS and form a network, a typical alkoxy condensation<br />

route is utilized by reacting the silanol-terminated PVMS with tetraethoxysilane (TEOS) as<br />

the cross-linker and a tin catalyst with a molar ratio of TEOS:PVMS:catalyst of 1:7:1 in<br />

order to produce films with the highest elasticity. The PVMS solution can be placed on glass<br />

slides and cured either at 60 °C or at room temperature for 30 minutes.<br />

Shell cross-linked micelles at a concentration of 5 mg/mL of hexane were added to the<br />

PVMS:TEOS mixture prior to catalyst addition at ratios of PFS-b-PI:PVMS of 1:150, 1:300,<br />

and 1:600 and cured at both ambient temperature and 60 °C for 30 minutes. Films were<br />

microtomed below the Tg of PVMS (-128 °C). After probing the interior of the PVMS films,<br />

we see the appearance of cylindrical micelles, spherical micelles, and iron clusters, as<br />

depicted in Figure 12. As far as the structure of any remaining micelles, there is not an<br />

appreciable difference of the morphology in the ambient vs. 60°C cured films. This<br />

demonstrates that the micelles are not affected by the curing temperature, which is well<br />

below the Tg. Lengths of cylindrical micelles range from 20-400 nm, which is a drastic<br />

reduction from the length of PFS-b-PI in solution which exceeds one micron. The reason for<br />

this is explored by exposing by heating and agitating the micelle solution at conditions that<br />

imitate the crosslinking reaction conditions.<br />

163


Shell-cross-linked PFS-b-PI were dropcast from hexane onto a carbon coated TEM grid<br />

were found to have an average length of 810 ± 300 nm. Some affinity exists between<br />

micelles both in an end-to-end and side-to-side fashion. When the solution is hand-agitated,<br />

similar to the conditions by which its necessary to mix the catalyst into the PVMS solution,<br />

we see that the average length is decreased to 620 ± 290 nm. Likewise less affinity exists<br />

between the micelles. Some of the cylindrical micelles also collapse into spherical micelles,<br />

as seen in Figure 5.13b. When the solution is heated in an oven at 60 °C for 15 minutes, we<br />

see little change in the average length (840 ± 380). However the largest change is that we see<br />

networks of micelles formed and much more curvature. These results are indicative that<br />

heating the micelle solution prior to cross-linking may achieve a more continuous network<br />

throughout the PVMS. These results are summed up in Figure 5.13.<br />

Complementary SEM images were also taken on both the surface of the PVMS films and<br />

also the cross-sections. Of interest primarily was the 1:150 phase-separated film. An image<br />

of the the surface shows large protrusions from the film, edx reveals these are largely<br />

composed of silicon. Due to the fact that the PVMS solution was not stirred prior to cross-<br />

linking, it is not unusual that some parts of the solution cross-linked quicker before the<br />

matrix did as a whole and exist as flocs in the matrix. These parts are protruding from the<br />

surface of the film. In addition, cross-sections of these films were taken by exposing the<br />

films to liquid nitrogen and then immediately fracturing. Figure 5.14 shows the SEM<br />

micrographs of changes in surface density as well as the fractured surface.<br />

164


5.5. Conclusions<br />

In this work PFS is investigated as electrospun fibers and PFS-b-PI cylindrical micelles<br />

are cross-linked within an elastomeric PVMS matrix. In order to produce uniform, beadles<br />

fibers from electrospinning it is necessary to include the surfactant Triton-X in order to<br />

reduce the surface tension in the solution. Electrospinnning PFDMS demonstrates an increase<br />

in interplanar distance between parallel ferrocene chains due to the elongational forces during<br />

electrospinning. Adding in small iron oxide nanoparticles (~10 nm) further increases this<br />

interplanar distance since the polymer chains orient around the nanoparticles. In addition,<br />

although the completely polymer PFMPS is completely amorphous, we see electrospining-<br />

induced crystallinity. Cylindrical micelles of PFS-b-PI can be shell cross-linked and<br />

successfully maintain their shape when cross-linked in a PVMS matrix. Although the length<br />

sof these molecules decrease from greater than 1.5 microns in solution to less than 400 nm<br />

when dispersed in the PVMS matrix this demonstrates a great starting point for soft nanowire<br />

devices utilizing the crystalline core of PFS block copolymers.<br />

165


Tables<br />

Table 5.1: Characteristics of the (Co)Polymers Used in this Study<br />

Polymer Mn (kDa) PDI ΦPFS<br />

Poly(ferrocenyldimethylsilane) 95.6 1.42 1.00<br />

Poly(ferrocenylmethylphenylsilane) 109.6 1.12 1.00<br />

Poly(ferrocenylethylmethylsilane) 38.0 1.10 1.00<br />

Poly(2-vinylpyridine)-bpoly(ferrocenylethylmethylsilane)<br />

20.0 1.17 0.20<br />

Poly(2-vinylpyridine)-bpoly(ferrocenydimethylsilane)<br />

20.0 1.16 0.15<br />

Poly(ferrocenyldimethylsilane)-bpoly(isoprene)<br />

32.0 1.03 0.14<br />

Poly(ferrocenyldimethylsilane)-bpoly(isoprene)<br />

(cross-linked corona)<br />

56.8 1.11 0.09<br />

Poly(2-vinyl pyridine) 78.5 — —<br />

Poly(isoprene) 80.0 1.01 —<br />

Poly(ε-caprolactone) 80.0 — —<br />

Poly(vinyl methoxysilane) 30.0 — —<br />

166


Polymer Solvent<br />

15%<br />

PFDMS<br />

15%<br />

PFDMS<br />

20%<br />

PFDMS<br />

18%<br />

PFPMS<br />

Table 5.2: PFDMS Fiber Diameters<br />

9:1 v:v<br />

THF:DMF<br />

9:1 v:v<br />

THF:DMF<br />

167<br />

2 vol %<br />

Surfactant<br />

Mean<br />

Diameter (nm)<br />

St Dev<br />

Yes 1320 980<br />

No 1800 400<br />

Dichloromethane Yes 610 220<br />

9:1 v:v<br />

THF:DMF<br />

Yes 1260 540


Table 5.3: X-ray diffraction bragg angle, d-spacing, and average crystallite size for 96 kDa<br />

poly(ferryocenydimethylsilane) powder and electrospun fibers<br />

PFDMS Powder<br />

PFDMS Fiber<br />

PFDMS Fiber<br />

+ 14nm NP<br />

PFDMS Fiber<br />

+ 10nm NP<br />

Angle<br />

(°2Ɵ)<br />

168<br />

d-spacing<br />

(nm)<br />

��(nm)�<br />

13.95 0.63 5.74<br />

20.09 0.44 19.90<br />

23.39 0.38 6.69<br />

13.55 0.65 17.20<br />

18.55 0.47 16.10<br />

25.62 0.35 2.70<br />

13.62 0.65 1.69<br />

23.43 0.37 30.50<br />

13.55 6.53 2.47


Table 5.4: Components utilized in the vulcanization of poly(isoprene)<br />

169


Table 5.5: Solubility parameters of polymers relative to the solvent n-hexane.<br />

Component<br />

170<br />

Solubility<br />

Parameter (δ)<br />

PI 16.5<br />

PVMS 15.1<br />

Hexane 14.9


Figures<br />

Fe<br />

tmeda<br />

Fe<br />

BuLi<br />

Fe<br />

a.<br />

Si + BuLi Fe<br />

Li<br />

. tmeda +<br />

Li<br />

Si<br />

Li<br />

Bu<br />

+ Si<br />

Fe<br />

Figure 5.1: Schematic of PFEMS synthesis in which the noted molecules are a)<br />

ferrocenophane b) ethylmethylsilaferrocenophane and c) poly (ferrocenylethylmethylsilane).<br />

171<br />

Si<br />

Cl<br />

Cl<br />

2 LiCl<br />

BuLi<br />

Fe<br />

Fe<br />

c.<br />

Si<br />

b.<br />

Si<br />

n


a.<br />

c.<br />

b.<br />

d.<br />

Figure 5.2. SEM micrographs of a) 15% PFDMS in THF:DMF without surfactant b) 15%<br />

PFDMS in THF:DMF with surfactant c) 20% PFDMS in DCM with surfactant and d) 18%<br />

PFPMS in THF:DMF with surfactant. All scale bars represent 20 μm.<br />

172


a<br />

.<br />

c<br />

b.<br />

Figure 5.3. PS-b-PFS-b-P2VP lithographic template used for the preparation of Nanoscale<br />

magnetic dots. a) Phase-separation of the triblock in the bulk. b) Hollow PFS cylinders are<br />

formed after etching because it has a selective resistance. c) Profile of the hollow PFS<br />

cylinders. (Used with permission by Jessica Gwyther from BristolUniversity.)<br />

173<br />

PFS hollow cylinders<br />

Etched<br />

P2VP<br />

matrix<br />

Etched<br />

PS<br />

holes


10 15 20 25<br />

2�<br />

Figure 5.4. XRD curves from 2θ = 7 – 25 °for PFDMS powder, fibers, and fibers with both<br />

10 and 14 nm iron oxide nanoparticles.<br />

174<br />

Fiber + 14 nm NP<br />

Fiber + 10 nm NP<br />

PFDMS Fiber<br />

PFDMS Powder<br />

Intensity


a<br />

.<br />

b.<br />

c.<br />

d.<br />

Figure 5.5 Schematic demonstrating the chains of PFDMS and corresponding d-spacing<br />

between adjacent ferrocene units in a) powder form, b) in electrospun fibers, c) in<br />

electrospun fibers with larger (~14 nm) iron oxide nanoparticles, and d) in electrospun fibers<br />

with smaller (~10 nm) iron oxide nanoparticles.<br />

175


Figure 5.6. TEM micrographs of PFS-b-PI micelles dropcast from a 1 mg/mL hexane<br />

solutiononto a carbon-coated TEM grid with an average width of 14.9 nm and lengths<br />

exceeding one micron.<br />

176


a.<br />

Figure 5.7. TEM micrographs of PFS-b-PI micelles blended at a ratio of a) 3:100 with PI and<br />

b) 3:25 with PI.<br />

177


Add micelles and apply<br />

heat<br />

Figure 5.8. Schematic of the vulcanization and micelle crosslinking technique.<br />

178


Figure 5.9. TEM micrographs of 1:1000 PFS-b-PI:PI vulcanized at ~120 °C for 5 hours<br />

demonstrating a complete dissolution of the micelles with only iron nanoparticles remaining.<br />

179


Figure 5.10. TEM micrographs of shell cross-linked PFS-b-PI with an average width of 43<br />

nm and lengths exceeding 1.5 microns dropcast from a hexane solution.<br />

180


Figure 5.11. TEM micrographs of PFS-b-PI micelles blended with a 5 wt% PI solution in<br />

hexane and dropcast onto a carbon-coated TEM grid.<br />

181


Figure 5.12. TEM micrographs of microtomed PFS-b-PI micelles blended with PVMS at<br />

ratios of 1:300 and 1:600 at thicknesses of ~120 nm.<br />

182


a.<br />

c.<br />

b.<br />

Figure 5.13. TEM micrographs shell cross-linked PFS-b-PI dropcast from a hexane solution<br />

as a) a control, b) heated to 70 °C for 30 minutes and c) stirred for several minutes<br />

mimicking conditions during the cross-linking of the PVMS solution.<br />

183


a.<br />

b.<br />

Figure 5.14. SEM micrographs of a PVMS cross-linked film containing PFS-b-PI micelles a)<br />

looking down the surface from the cross-section and b) of the fractured cross-section where<br />

the scalebar of the inset refers to 200 μm.<br />

184


References<br />

(1) Nguyen P, G.-E. P., Manners I Chemistry Reviews 1999, 99, 1515.<br />

(2) Manners, I. Science 2001, 294, 1664.<br />

(3) Arimoto FS, H. A. Journal of the American Chemical Society 1955, 77, 6295.<br />

(4) Abd-el-Aziz, A. S.; Manners, I. Frontiers in Transition Metal-Containing Polymers; Wiley<br />

VCH: Hoboken, 2007.<br />

(5) Foucher DA, T. B.-Z., Manners I Journal of the American Chemical Society 1992, 114, 6246.<br />

(6) Manners, I. Candian Journal of Chemistry 1998, 76, 371; Resendes R, N. J., Fischer A, et al.<br />

Journal of the American Chemical Society 2001, 123, 2116.<br />

(7) Rulkens, R.; Lough, A. J.; Lovelace, S. R.et al. Journal of the American Chemical Society<br />

1996, 118, 12683.<br />

(8) Manners, I. Synthetic Metal-Containing Polymers; Wiley-VCH: Weinheim, 2004.<br />

(9) Kulbaba, K.; Manners, I. Macromolecular Rapid Communications 2001, 22, 711.<br />

(10) Hajduk, D. A.; Harper, P. E.; Gruner, S. M.et al. Macromolecules 1994, 27, 4063.<br />

(11) Yu, Y. S.; Eisenberg, A. Journal of the American Chemical Society 1997, 119, 8383.<br />

(12) Matsen, M. W. Macromolecules 1995, 28, 5765.<br />

(13) Horiuchi, S.; Fujita, T.; Hayakawa, T.et al. Langmuir 2003, 19, 2963.<br />

(14) IW, H. The Physics of Block Copolymers Oxford, UK, 1998.<br />

(15) Hadjichristidis, N.; Pitsikalis, M.; Iatrou, H.et al. Macromolecular Rapid Communications<br />

2003, 24, 979.<br />

(16) Ni, Y. Z.; Rulkens, R.; Manners, I. Journal of the American Chemical Society 1996, 118,<br />

4102.<br />

185


(17) Schubert, U. S.; Eschbaumer, C. Angewandte Chemie-International Edition 2002, 41, 2893.<br />

(18) Cao, L.; Manners, I.; Winnik, M. A. Macromolecules 2002, 35, 8258.<br />

(19) Wang, X. S.; Winnik, M. A.; Manners, I. Angewandte Chemie-International Edition 2004,<br />

43, 3703.<br />

(20) Rider, D. A.; Cavicchi, K. A.; Power-Billard, K. N.et al. Macromolecules 2005, 38, 6931.<br />

(21) Kim, K. T.; Vandermeulen, G. W. M.; Winnik, M. A.et al. Macromolecules 2005, 38, 4958.<br />

(22) Kloninger, C.; Rehahn, M. Macromolecules 2004, 37, 8319; Lammertink, R.; Hempenius,<br />

M.; Thomas, E.et al. Journal of Polymer Science: Part B: Polymer Physics 1998, 37, 1009; Temple,<br />

K.; Kulbaba, K.; Power-Billard, K. N. Advanced Materials 2003, 15, 297.<br />

(23) Lastella, S.; Jung, Y. J.; Yang, H. C.et al. Journal of Materials Chemistry 2004, 14, 1791; Lu,<br />

J. Q.; Kopley, T. E.; Moll, N.et al. Chemistry of Materials 2005, 17, 2227.<br />

(24) Temple, K.; Kulbaba, K.; Power-Billard, K. N.et al. Advanced Materials 2003, 15, 297.<br />

(25) Hinderling, C.; Keles, Y.; Stockli, T.et al. Advanced Materials 2004, 16, 876.<br />

(26) Massey, J. A.; Power, K. N.; Winnik, M. A.et al. Advanced Materials 1998, 10, 1559.<br />

(27) Raez, J.; Manners, I.; Winnik, M. A. Journal of the American Chemical Society 2002, 124,<br />

10381.<br />

(28) Raez, J.; Manners, I.; Winnik, M. A. Langmuir 2002, 18, 7229.<br />

(29) Massey JA, T. K., Cao L, et al. Journal of the American Chemical Society 2000, 122, 11577.<br />

(30) Wang, H.; Winnik, M. A.; Manners, I. Macromolecules 2007, 40, 3784.<br />

(31) Wang, X.; Liu, K.; Aresenault, A. C.et al. Journal of the American Chemical Society 2007,<br />

129, 5630.<br />

(32) Wang, X. S.; Wang, H.; Coombs, N.et al. Journal of the American Chemical Society 2005,<br />

127, 8924.<br />

186


(33) Bronstein, L. M.; Huang, X.; Retrum, J.et al. Chem. Mater. 2007, 19, 3624.<br />

(34) Bronstein, L. M.; Atkinson, J. E.; Malyutin, A. G.et al. Langmuir 2011, 27, 3044.<br />

(35) Shtykova, E. V.; Huang, X.; Remmes, N.et al. J. Phys. Chem. B 2007, 111, 18078.<br />

(36) Reneker, D. H.; Chun, I. Nanotechnology 1996, 7, 216.<br />

(37) Muller, K.; Quinn, J. F.; Johnston, A. P. R.et al. Chem. Mater. 2006, 18, 2397.<br />

(38) Zhang, D.; Karki, A. B.; Rutman, D.et al. Polymer 2009, 50, 4189.<br />

(39) Chen, Z.; Foster, M. D.; Zhou, W.et al. Macromolecules 2001, 34, 6156.<br />

(40) Lin, T.; Wang, H. X.; Wang, H. M.et al. Nanotechnology 2004, 15, 1375.<br />

(41) P, S. Gottinger Nachrichten Gesell. 1918, 2, 98.<br />

(42) Kakade, M. V.; Givens, S.; Gardner, K.et al. Journal of the American Chemical Society 2007,<br />

129, 2777.<br />

(43) Rulkens, R.; Lough, A. J.; Manners, I.et al. Journal of the American Chemical Society 1996,<br />

118, 12683.<br />

(44) Rasburn, J.; Petersen, R.; Jahr, T.et al. Chemistry of Materials 1995, 7, 871.<br />

(45) Lyatskaya, Y.; Gersappe, D.; Balazs, A. C. Macromolecules 1995, 28, 6278.<br />

(46) Muller, M.; Gompper, G. Physical Review E 2002, 66.<br />

(47) Bernard, B.; Brown, H. R.; Hawker, C. J.et al. Macromolecules 1999, 32, 6254.<br />

(48) Leibler, L. Makromolekulare Chemie-Macromolecular Symposia 1988, 16, 1.<br />

(49) Winey, K. I.; Thomas, E. L.; Fetters, L. J. Macromolecules 1992, 25, 2645.<br />

(50) Wang, X. S.; Arsenault, A.; Ozin, G. A.et al. Journal of the American Chemical Society<br />

2003, 125, 12686.<br />

(51) Wang, X. S.; Guerin, G.; Wang, H.et al. Science 2007, 317, 644.<br />

(52) Xia, Y. N.; Yang, P. D.; Sun, Y. G.et al. Advanced Materials 2003, 15, 353.<br />

187


(53) Morrison, N. J.; Porter, M. Rubber Chemistry and Technology 1984, 57, 63.<br />

(54) Picart, C.; Lavalle, P.; Hubert, P.et al. Langmuir 2001, 17, 7414.<br />

(55) Xie, R.; Karim, A.; Douglas, J. F.et al. Physical Review Letters 1998, 81, 1251.<br />

(56) Hanley, K. J.; Lodge, T. P.; Huang, C. I. Macromolecules 2000, 33, 5918.<br />

188


6.1 Conclusions<br />

CHAPTER VI<br />

Conclusions and Future Work<br />

In this dissertation, we determined how to controllably align and position metal-<br />

containing nanomaterials within electrospun polymer fibers, develop control over<br />

electrospinning parameters to create organic-inorganic interactions, and maximize<br />

functionalities of the organometallic filler materials. Firstly, we demonstrated the concept of<br />

hierarchical alignment of fibers containing aligned gold nanorods over macroscopic<br />

dimensions through electrospinning . We hypothesized that control over the location and<br />

alignment of nanoparticles within electrospun fibers can be achieved through the<br />

straightforward use of an external electromagnetic field applied during electrospinning.<br />

Likewise, thermodynamic incompatible polymer blends were prepared in order to control the<br />

spatial location of superparamagnetic iron oxide nanoparticles within electrospun fibers.<br />

Lastly, organometallic polymers were utilized to create functional nanocomposites that<br />

married the highly desirable properties of metals with those of polymers. This approach is<br />

one that can be applied to a wide variety of both filler materials and polymer matrix systems,<br />

which will broadly contribute to the field of material science.<br />

189


6.1.1 Long-Range Alignment of Gold Nanorods in Electrospun Polymer Nano/Microfibers<br />

The nanoscale orientation of GNRs was achieved with scalable, macroscopic order a<br />

distance of several centimeters. Here, GNRs with an aspect ratio of 3.1 exhibit excellent<br />

alignment with their longitudinal axes parallel to the fiber axis n for electrospun polymer<br />

nano/microfibers with diameters of 40-600 nm, and they maintain substantial alignment in<br />

microfibers measuring up to 3000 nm in diameter. Loading of GNRs can be varied with no<br />

discernible impact on the net degree of alignment while fiber diameter has a more direct<br />

correlation. Electron diffraction measurements of the aligned GNRs confirm preferred<br />

orientation of the {100} and {111} GNR planes. Optical absorbance spectroscopy<br />

measurements performed on macroscopically aligned electrospun fibers containing aligned<br />

GNRs demonstrate that the longitudinal surface plasmon resonance bands are polarization<br />

dependent and display maximum absorption when the polarizer is parallel to n.<br />

6.1.2 Magnetic Field-Induced Alignment of Nanoparticles in Electrospun Microfibers<br />

Nanoscale alignment of superparamagnetic iron oxide nanoparticles (SPIONs) with<br />

diameters ranging from 12 to 18 nm has been achieved through the coupling of an external<br />

magnetic field with the electric field in the electrospinning process. Microfibers containing<br />

one-dimensional arrays of SPIONs extending beyond one micron in length result from an<br />

perpendicular electromagnetic field of 26 mT. In the present system investigated, a SPION<br />

concentration of greater than 0.05 vol% is required to induce discernible alignment.<br />

Moreover, alignment of the electrospun microfibers through the use of various established<br />

methods can yield nanocomposites with multiscale (i.e., nanoscale and macroscale)<br />

190


anisotropic properties. Complementary superconducting quantum interference device<br />

(SQUID) measurements reveal that the saturation magnetization is significantly lower for<br />

SPIONs in electrospun fibers with or without magnetic field-induced alignment than for<br />

unembedded SPIONs. The mean magnetic moment increases with improving alignment,<br />

however,, which demonstrates that nanoscale alignment of SPIONs affects their intrinsic<br />

physical properties.<br />

6.1.3 Using Polymer Blend Morphology to Position Ligand-Functionalized Nanoparticles<br />

in Electrospun Polymer Microfibers<br />

Blends of hydrophobic (P2VP) and hydrophilic polymers (PEO) have been prepared<br />

to discern the feasibility of controlling the spatial location of SPIONs within electrospun<br />

fibers on the basis of thermodynamic compatibility. In this case, a core-sheath structure<br />

naturally forms with the hydrophobic SPIONs sequestered in one preferential phase at 30<br />

wt% PEO. For both a pure PEO solution and a pure P2VP solution a decrease in zero shear<br />

viscosity was found upon the addition of SPIONs. Interfacial distances ranged from 26 – 109<br />

nm demonstrating diffuse interfaces between the two immiscible polymers. X-ray diffraction<br />

(XRD) confirms that PEO crystals becomes more aligned upon SPION addition suggesting<br />

that the nanoparticles may reside at the interface between the PEO and P2VP phases.<br />

6.1.4. Nanostructured Organometallic Polymer Systems Containing Poly(ferrocenylsilanes)<br />

Lastly, poly(ferrocenylsilanes) (PFS) are investigated as the fiber-forming polymer in<br />

an electrospinning system. Uniform fibers are proven to be formed from the addition of a<br />

surfactant to a solution of polyferrodimethylsilane (PFDMS) in dichloromethane.<br />

191


Electrospun PFDMS fibers demonstrates an increase in interplanar distance between parallel<br />

ferrocene chains due to the elongational forces during electrospinning when compared to the<br />

pure powder. The addition of small iron oxide nanoparticles (~10 nm) further increases this<br />

interplanar distance where we hypothesize that the polymer chains orient around the<br />

nanoparticles. In addition, although the polymer poly(ferromethylphenylsilane) (PFMPS) is<br />

completely amorphous, we see electrospining-induced crystallinity. Cylindrical micelles of<br />

PFS-b-PI can be shell cross-linked and successfully maintain their shape when cross-linked<br />

in a poly(vinyl methoxysilane) (PVMS) matrix. Although the length sof these molecules<br />

decrease from greater than 1.5 microns in solution to less than 400 nm when dispersed in the<br />

PVMS matrix this demonstrates a great starting point for soft nanowire devices utilizing the<br />

crystalline core of PFS block copolymers.<br />

6.2 Recommendations for future work<br />

6.2.1 Gold Nanorod Alignment through Electrospun Fiber Degradation<br />

Utilization of polymer matrices to align gold nanorods can produce macroscale<br />

alignment but a method to self-assemble gold nanorods in solution or on a substrate without<br />

utilizing a hierarchical structure is needed. In this work the average deviant angle between<br />

the GNR axis and the fiber axis for small fibers (


a 35-46% decrease in molecular weight. 2 This same principal could be applied to PEO<br />

nanofibers at various wavelengths and times to see if it would be possible to completely<br />

degrade the PEO matrix. This would require aligned fibers with very small fiber diameters;<br />

in addition another variable that could be explored is the effect UV exposure on a monolayer<br />

of fibers versus multilayers of fibers. Optical absorbance spectra could be collected at<br />

various times of UV exposure to remove the troublesome polymer scattering in the spectra<br />

and provide a more precise way to monitor GNR alignment. Dissolution of PEO by water or<br />

a solvent such as chloroform could also be attempted to see if aligned GNRs could result. If<br />

either of these techniques were successful it could be possible to align GNRs onto a substrate<br />

used for biomedical imaging. 3<br />

6.2.2 Field Uniformity in Magnetic-Assisted Electrospinning<br />

In this work, alignment of SPIONs in electrospun fibers was achieved through the use<br />

of an external u-shaped electromagnet in conjunction with the electric field during<br />

electrospinning. However, alignment was not perfect due to several imperfections; namely,<br />

whipping of the jet and a lack of magnetic field uniformity due to the shape of the<br />

electromagnet used. An electromagnet utilizing a hollow cylinder, wrapped in magnetic wire<br />

connected to a voltage source would provide a much more uniform field and allow the<br />

spinneret needle to be placed ‘inside’ the magnet. In order to achieve a magnetic field<br />

equivalent to that of the electromagnet, 26 mT, hundreds of coils would have to be placed<br />

over the hollow cylinder. Another follow-up experiment for this work would be to determine<br />

a way to decrease the interparticular distance between SPIONs in the electrospun fibers so<br />

193


that they act as a single ‘nanowire’ and not as individual particles. This could lead to some<br />

very interesting superparamagnetic properties; such as a larger remanence magnetization and<br />

coercivity for aligned samples. 4 This could be investigated by using particles that are<br />

surface-functionalized with a different ligand, by altering the length of the ligand on the<br />

nanoparticle surface, or look at another type of magnetic nanoparticle such as cobalt<br />

nanoparticles. 5<br />

6.2.3 Poly(ferrocenylsilane) Cylindrical Micelles Oriented within Electropun Fibers<br />

The focus of this thesis has been on controlling the spatial location of nanoscale<br />

objects within electrospun fibers in order to create functional nanocomposites. A natural<br />

extension would be to spin PFS block copolymer cylindrical micelles to see if they orient<br />

within electorspun fibers and investigate the resulting properties. Self-assembly of PFS-b-<br />

poly(2-vinyl pyridine) (P2VP) has been proven to form cylindrical micelles in alcohol<br />

solvents 6 such as ethanol and isopropanol. If these micelles could be formed in a solvent<br />

with a higher dielectric constant, such as dimethylformamide (DMF), then it would be<br />

suitable for electrospinning. The self-assembled cylinders could then be blended with P2VP<br />

to explore the interactions between a diblock and homopolymer binary blend behave under<br />

longitudinal forces during electrospinning. The time scales at which self-assembly occurs<br />

can be investigated by electrospinning both assembled and unassembled micelles. If the<br />

assembled micelles do in fact stay intact during electrospinning, this will be an indication of<br />

their robustness and also may stabilize the electrospun fibers much like carbon nanotubes 7 .<br />

194


The location of the micelle within the fiber (on the surface, random, oriented) will also give<br />

information about the diblock/hompolymer interactions.<br />

PFDMS-b-P2VP, at a concentration of 2 mg/mL, was dissolved in DMF, a P2VP<br />

selective solvent, and after 24 hours formed cylinders with diameters ranging from 100 -<br />

1000 nm, which can be seen in the TEM images in Figure 6.1. This has not been previously<br />

reported in literature. P2VP (200 kDa) was electrospun from DMF at initial concentrations<br />

ranging from 14-20%. At lower concentrations, very thin fibers were formed but with a large<br />

proportion of beads. Electrospinning at 20% produced thin fibers, with a mean diameter of<br />

260 ± 170 nm, and few bead defects. To ensure this criteria was met, instead of increasing<br />

the concentration of P2VP, which has been shown to reduce bead defects 8 , the solvent was<br />

changed to volume ratio of 9:1 DMF:ethanol. The addition of ethanol reduces the surface<br />

tension of the jet and thus also reduces bead defects. Initial concentrations of 21% P2VP<br />

prepared in a 9:1 DMF:ethanol solution was blended with 0.8 mg of PFDMS-b-P2VP in<br />

DMF. The solution was gently shaken to disperse the two systems. The addition of the<br />

micelles did not disturb the solution during electrospinning, fibers without beads were<br />

formed with a mean diameter of 300 ± 190 nm, very similar to the pure P2VP solution.<br />

Energy-dispersive x-ray spectroscopy (EDX) results confirmed that no iron was present on<br />

the surface and scanning electron microscopy (SEM) revealed that the fiber surfaces did not<br />

appear different before and after micelle addition (Figure 6.2). It was not possible to see the<br />

micelles within the fibers through transmission electron microscopy (TEM), even when the<br />

fibers were very small and more transparent to the electron beam as seen in Figure 6.3. To<br />

195


emedy this, the electrospun fibers were aligned and coated with 5 nm of Pt/Pd, and then<br />

cured within an epoxy matrix for microtoming. Fiber cross-sections of ~120 nm were cut on<br />

a room temperature microtome and the internal structure was probed via TEM. It appears as<br />

if the micelles have stayed intact during the electrospinning process, (Figure 6.4), however<br />

the structures depicted are not oriented parallel to the fiber axis (since cross-sections would<br />

only present the top face of the cylinders, or just a sphere). The micelles are likely<br />

orthogonal to the fiber axis. In order to probe this result further, higher concentrations of<br />

micelles should be included in the solution to be electrospun and very small fiber diameters<br />

should be prepared. It would then be beneficial to microtome perpendicular to the fiber<br />

surface, since we would expect the cylinder-like rods to be aligned to the fiber axis, as is the<br />

case with CNTs. 9 In conclusion, extending the topic of this dissertation to the case of block<br />

copolymer micelles within electrospun fibers would be an excellent inclusion and would<br />

further probe the dynamics between a diblock copolymer/hompolymer blend in an<br />

electrospun fiber.<br />

196


Figures<br />

Figure 6.1. Dropcast PFDMS-b-P2VP cylindrical micelles onto a carbon coated TEM grid<br />

from DMF.<br />

197


a.<br />

b.<br />

Figure 6.2. SEM micrographs of 32 wt% P2VP fibers electrospun from 9:1 DMF:THF a)<br />

without micelle addition and b) with PFDMS-b-P2VP micelles.<br />

198


Figure 6.3. TEM micrograph of 32 wt% P2VP fibers electrospun with PFDMS-b-<br />

P2VP micelles.<br />

199


Figure 6.4. TEM micrographs of microtomed P2VP fibers containing PFDMS-b-P2VP<br />

micelles.<br />

200


References<br />

(1) van der Zande, B. M. I.; Pagès, L.; Hikmet, R. A. M.et al. J. Phys. Chem. B 1999, 103, 5761;<br />

Hore, M. J. A.; Composto, R. J. ACS Nano 2010, 4, 6941.<br />

(2) Dong, Y. X.; Yong, T.; Liao, S.et al. Tissue Engineering Part A 2008, 14, 1321.<br />

(3) Durr, N. J.; Larson, T.; Smith, D. K.et al. Nano Lett. 2007, 7, 941.<br />

(4) Park, J. I.; Jun, Y. W.; Choi, J. S.et al. Chem. Commun. 2007, 5001.<br />

(5) Lu, A. H.; Salabas, E. L.; Schuth, F. Angewandte Chemie-International Edition 2007, 46,<br />

1222.<br />

(6) Wang, X.; Liu, K.; Aresenault, A. C.et al. Journal of the American Chemical Society 2007,<br />

129, 5630; Wang, H.; Winnik, M. A.; Manners, I. Macromolecules 2007, 40, 3784.<br />

(7) Ko, F.; Gogotsi, Y.; Ali, A.et al. Advanced Materials 2003, 15, 1161.<br />

(8) Fong, H.; Chun, I.; Reneker, D. H. Polymer 1999, 40, 4585.<br />

(9) Dror, Y.; Salalha, W.; Khalfin, R. L.et al. Langmuir 2003, 19, 7012.<br />

201


APPENDIX<br />

202


APPENDIX I<br />

Responsive PET Nano/Microfibers via Surface-Initiated Polymerization<br />

Abstract<br />

A. Evren Özçam, Kristen E. Roskov, Jan Genzer and Richard J. Spontak<br />

Poly(ethylene terephthalate) (PET) is one of the most important thermoplastics in<br />

ubiquitous use today due to its mechanical properties, clarity, solvent resistance and<br />

recyclability. In this work, we functionalize the surface of electrospun PET microfibers by<br />

growing poly(N-isopropylacrylamide) (PNIPAAm) brushes through a chemical sequence that<br />

avoids PET degradation to generate thermoresponsive microfibers that remain mechanically<br />

robust. Amidation of deposited 3-aminopropyltriethoxysilane, followed by hydrolysis, yields<br />

silanol groups that permit surface attachment of initiator molecules, which can be used to<br />

grow PNIPAAm via "grafting from" atom-transfer radical polymerization. Spectroscopic<br />

analyses performed after each step confirm the expected reaction and the ultimate growth of<br />

PNIPAAm brushes. Water contact-angle measurements conducted at temperatures below and<br />

above the lower critical solution temperature of PNIPAAm, coupled with adsorption of Au<br />

nanoparticles from aqueous suspension, demonstrate that the brushes retain their reversible<br />

thermoresponsive nature, thereby making PNIPAAm-functionalized PET microfibers<br />

suitable for filtration media, tissue scaffolds, delivery vehicles, and sensors requiring robust<br />

microfibers.<br />

203


Introduction<br />

Electrospinning is an emerging fabrication technique capable of generating solid polymer<br />

fibers that range from tens of nanometers to several microns in diameter. Such<br />

nano/microfibers are of fundamental and technological interest due to their high surface-to-<br />

volume ratio. During wet electrospinning, a polymer solution of sufficiently high viscosity<br />

and conductivity is subjected to an electric field. When the electrostatic forces overcome<br />

surface tension, a charged jet emitted from the tip of a Taylor cone 1 undergoes a whipping<br />

action 2 wherein the solvent evaporates, and is subsequently collected as a dry, randomly<br />

oriented fiber mat on a grounded collector plate. This process strategy is appealing due to the<br />

simple setup required and the ability to tailor fiber characteristics with relative ease. 3<br />

Although the morphology of electrospun nano/ microfibers is desirable, they tend to lack the<br />

functionality that is sought in contemporary applications. One way to overcome this<br />

deficiency is by developing multicomponent nano/ microfibers, in which the fiber-forming<br />

polymer is modified with one or more species designed to enhance targeted properties. 4<br />

Surface-active compounds added to the polymer solution prior to electrospinning may,<br />

however, remain trapped within the resultant fiber upon solidification and thus exhibit<br />

substantially reduced activity. 5 While antibacterial biocides incorporated in this fashion lose<br />

much of their efficacy, 6 quaternary ammonium species covalently bonded to as-spun fibers<br />

can create a permanent antibacterial surface. 7 Alternatively, polarizable antibacterial<br />

copolymers co-dissolved with the fiber-forming polymer can be brought to the fiber surface,<br />

where they remain anchored in place, by the electric field during electrospinning. 8 Recently,<br />

204


Agarwal et al. 9 have surveyed chemical routes by which to modify and functionalize the<br />

surface of electrospun nanofibers for diverse applications ranging from functional textiles,<br />

catalyst supports and ion-exchange membranes to drug delivery and tissue engineering.<br />

Polymers such as poly(ethylene terephthalate) (PET), which is widely known for its<br />

mechanical strength, transparency and solvent resistance, tend to possess a hydrophobic<br />

surface and a low surface energy, 10 in which case electrospun nano/microfibers require post-<br />

treatment so that chemically-active species are positioned on the fiber surface. Methods by<br />

which to achieve such surface functionalization include UV treatment, 6 mineralization, 11<br />

core-shell formation, 12 chemical vapor deposition 13 or inclusion of reactive compounds. 14<br />

Once these chemically-active groups are available, covalent bonding, 15 immobilization 16 or<br />

electrostatic interactions 17 can be used to introduce functional moieties to the fiber surface<br />

without adversely affecting the bulk fiber properties. While surface modification could<br />

permit the use of electrospun PET 18 nano/microfibers in filtration media, 19 protective<br />

textiles, 20 tissue scaffolds, 21 and drug-delivery vehicles, 22 most of the modification<br />

approaches listed above purposefully or inadvertently promote PET degradation. Thus, the<br />

conditions by which surface modification is conducted must be monitored carefully to avoid<br />

compromising the bulk properties of PET.<br />

Grafting polymer brushes represents an alternative approach by which to modify and<br />

control the surface properties of materials. 23 Numerous studies reported on surface-initiated<br />

grafting on surfaces of various geometries with a plethora of different monomers by<br />

employing numerous polymerization routes. Poly(N-isopropylacrylamide) (PNIPAAm) is<br />

205


solely considered because of its thermoresponsive nature 24 (it possesses a lower critical<br />

solution temperature, Tc, in water at ≈32°C). Prior efforts to polymerize styrene 25 and<br />

NIPAAm 26 on flat PET surfaces have relied on different means of activating the PET surface<br />

(e.g., saponification, plasma treatment and aminolysis) for the purpose of attaching initiators.<br />

The major drawback of such treatments, however, is that they may increase surface<br />

roughness by degrading PET, which is of concern with regard to electrospun PET<br />

nano/microfibers. Independent studies 27 have confirmed that 3-aminopropyltriethoxysilane<br />

(APTES) can be used to functionalize the surface of PET via amidation with negligible<br />

degradation of the parent PET material. Unlike short alkyl amines (which can diffuse into<br />

and react throughout, and thus weaken, PET 28 ), the bulky triethoxysilane group on APTES<br />

hinders diffusion, changes its chemical nature upon amidation and creates a barrier by<br />

restricting the diffusion of other APTES molecules. Moreover, since the ethoxysilane groups<br />

of APTES are exposed at the polymer/air interface, hydrolysis of triethoxysilane yields<br />

silanol groups that facilitate initiator attachment.<br />

Thermoresponsive PNIPAAm brushes on electrospun fibers have been recently reported.<br />

For instance, Brandl et al. 29 describe the synthesis of a copolymer of 2-hydroxyethyl<br />

methacrylate (HEMA) and methyl methacrylate (MMA) and its post-polymerization<br />

modification with 2-bromoisobutyrylbromide to prepare a macroinitiator. They claim that<br />

electrospinning of the macroinitiator and subsequent polymerization of NIPAAm results in<br />

thermoresponsive polymer brushes. The disadvantage of this technique is that the location of<br />

the "active" initiator group in the fiber depends on the dielectrophoretic forces, polarizability<br />

206


contrast and surface tension of the comonomers, which invariably reduces the concentration<br />

of "active" initiator centers on the fiber surface. The presence of ester groups between the<br />

butyrylbromide group of the initiator and hydroxyethyl group of HEMA likewise yields<br />

hydrolyticly unstable bonds at pH values greater than 8 and lower then 5. Furthermore, the<br />

presence of HEMA comonomer on the macroiniatitor may result in swelling and absorption<br />

of NIPAAm monomer by the electrospun fiber in the aqueous polymerization medium.<br />

Similarly, Fu et al. 30 have synthesized and electrospun a copolymer of 4-vinylbenzylchloride<br />

and glycidyl methacrylate. Subsequent modification of the electrospun fibers with sodium<br />

azide produces azide surface groups that can be coupled with alkyne-functionalized<br />

PNIPAAm chains via a click reaction to generate PNIPAAm surface chains that affect the<br />

wettability of the fibers. Such grafting of PNIPAAm brushes on electrospun fibers would be<br />

necessarily low because of the sparse population of active azide groups on the fiber surface<br />

and the accompanying steric hindrance caused by the "grafting to" polymerization technique.<br />

In this work, we report a robust and universal way of preparing functional and<br />

thermoresponsive PNIPAAm brushes that are covalently attached to electrospun PET fibers.<br />

First, electrospun PET microfibers are modified with APTES to create surface-bound<br />

hydroxyl groups for the attachment of [11-(2-bromo-2-methyl)propionyloxy]<br />

undecyltrichlorosilane (BMPUS), which serves as an ATRP initiator for the polymerization<br />

of NIPAAm. Several analytical techniques are employed to (i) characterize the properties of<br />

as-spun and post-modified PET microfibers and (ii) follow the polymerization of NIPAAm<br />

via ATRP. In addition, we investigate the thermoresponsive nature of PNIPAAm-decorated<br />

207


PET microfibers by attaching Au nanoparticles at temperatures above and below the Tc of<br />

PNIPAAm.<br />

Experimental<br />

Food-grade recycled PET flakes were kindly supplied by the United Resource Recovery<br />

Corp. (Spartanburg, SC). The HFIP was obtained from Oakwood Products Inc. (Estill, SC),<br />

and anhydrous toluene, 2-chlorophenol, APTES, NIPAAm, copper I bromide (CuBr), copper<br />

II bromide (CuBr2), and N,N,N',N',N"-pentamethyldiethylenetriamine (PMDETA) were all<br />

purchased from Sigma-Aldrich and used as-received. The citrate-stabilized Au<br />

nanoparticles 29 (diameter = 16.9±1.8 nm) and BMPUS initiator 30 were prepared as described<br />

earlier. The PET flakes were dissolved in HFIP at different concentrations and electrospun at<br />

ambient temperature and 10 kV to generate microfibers varying in diameter. Thin films of<br />

PET measuring 12 and 180 nm thick, as discerned by ellipsometry (v.i.), were spun-cast at<br />

25ºC on silicon wafers from 0.5 and 3.0% (w/w) solutions, respectively, in 2-chlorophenol.<br />

Microfiber mats and thin films were stored under vacuum for at least 48 h prior to use to<br />

remove entrapped solvent.<br />

The APTES was deposited on the PET microfibers and thin films by exposing the<br />

samples to 1% (v/v) APTES/anhydrous toluene solutions for 24 h at ambient temperature,<br />

followed by sonication in toluene for 10 min to remove loosely adsorbed APTES molecules.<br />

The ethoxysilane groups of the surface-anchored APTES molecules were hydrolyzed in<br />

acidic water (pH ≈ 4.5-5.0) for 6 h at ambient temperature, and then the fiber mats were<br />

washed with copious amount of water. After drying the samples under reduced pressure,<br />

208


BMPUS was deposited on the PET-SiOH surfaces by established protocols. 31 The PNIPAAm<br />

brushes were subsequently grown from PET-SiOH surfaces by ATRP of NIPAAm, as<br />

described elsewhere. 29 Briefly, 6.30 g NIPAAm was dissolved in a mixture of 4.86 g<br />

methanol and 6.30 g water in an argon-purged Schlenk flask, and oxygen was removed via<br />

three freeze-thaw cycles. After removal of oxygen, PMDETA (0.56 g), CuBr (0.16 g) and<br />

CuBr2 (0.016 g) were added to the solution prior to an additional freeze-thaw cycle. The<br />

Schlenk flask was tightly sealed and transferred to an argon-purged glove box. Microfiber<br />

mats and thin films of PET were submersed in the solution for 8 h at ambient temperature,<br />

after which they were removed, promptly rinsed with methanol and deionized water, and<br />

then sonicated in deionized water for 10 min.<br />

The thickness of the thin PET films was measured by variable-angle spectroscopic<br />

ellipsometry (J.A. Woollam) at a 70º incidence angle before and after each modification step<br />

to discern the PNIPAAm brush height. Surface chemical composition was monitored by XPS<br />

performed on a Kratos Analytical AXIS ULTRA spectrometer at a take-off angle of 90º. The<br />

FTIR analysis of the PET microfibers was conducted in transmission mode on a Nicolet 6700<br />

spectrometer after embedding the microfiber mats in potassium bromide pellets. For each<br />

sample, 1024 scans were acquired after background correction at a resolution of 4 cm -1 .<br />

Resultant XPS and FTIR spectra were analyzed using the Vision and Omnic Specta software<br />

suites, respectively. The thermoresponsive behavior of PET and PET-PNIPAAm microfibers<br />

was interrogated by measuring the WCA at different temperatures via the sessile drop<br />

technique on a Ramé-Hart Model 100-00 instrument. As-spun and modified PET microfibers<br />

209


were coated with ≈16 nm of Au, and their diameter and surface morphology were examined<br />

by field-emission SEM performed on a JEOL 6400F electron microscope operated at 5 kV.<br />

Results and Discussion<br />

The diameters of electrospun PET microfibers, prepared according to the protocol<br />

provided in the Experimental section and measured by scanning electron microscopy (SEM),<br />

are 450±100, 800±200 and 1200±300 nm for 6, 8 and 10% (w/w) solutions, respectively, of<br />

PET in hexafluoroisopropanol (HFIP). The surfaces of unmodified PET microfibers<br />

consistently appear smooth (cf. Figure 1) with some slight dimpling observed occasionally<br />

along the fiber axis. Microfibers modified with thermoresponsive PNIPAAm brushes have<br />

been generated in a sequence of four steps, which are depicted schematically in Figure 1.<br />

Briefly, APTES molecules are attached to the PET surface via aminolysis between PET and<br />

the primary amine of APTES. Next, the ethoxysilane groups on APTES are hydrolyzed to<br />

generate silanol groups for BMPUS attachment. Finally, PNIPAAm brushes are grown<br />

directly from the PET microfiber surface. A second SEM image displaying PET microfibers<br />

modified with PNIPAAm brushes is included for comparison in Figure 1 to demonstrate that<br />

these microfibers appear marginally rougher than the as-spun microfibers at the end of the<br />

modification and brush growth process. The difference in microfiber morphology is almost<br />

indiscernible and the PNIPAAm brushes on spin-coated PET films on silicon wafers appears<br />

smooth, combination of these verifies that the brush is uniformly distributed on the surface of<br />

the microfibers. Below, we provide a detailed assessment of each of the steps in this<br />

polymerization sequence.<br />

210


In Figure 2, Fourier-transform infrared (FTIR) spectra are presented for three materials:<br />

(a) as-spun microfibers (PET), (b) APTES-modified PET microfibers following hydrolysis<br />

(PET-SiOH) and (c) PET microfibers with PNIPAAm brushes (PET-PNIPAAm). The<br />

appearance of new peaks located at 1650 cm -1 (amide I band) 1550 cm -1 (amide II band),<br />

1470 cm -1 , and 3300 cm -1 in Figure 2b is due to the formation of secondary amide groups,<br />

thereby confirming the grafting of APTES to the PET microfiber surface. Previous reports 27<br />

regarding the surface modification of PET with APTES could not detect the amidation<br />

reaction via FTIR due to a very low signal-to-noise ratio. Detection of these groups by FTIR<br />

in the present work is attributed to the large surface area afforded by the microfibers.<br />

Successful attachment of APTES can also be inferred from the surface properties of modified<br />

microfibers upon exposure to acidic water, which promotes hydrolysis of the ethoxysilane<br />

groups to silanol groups. Resulting changes in static water contact angle (WCA) and<br />

specimen thickness are measured on flat PET films spun-cast on silicon wafer.<br />

Corresponding values of WCA for films of PET-SiOH and PET after hydrolysis are 50�0.8˚<br />

and 71�0.8˚, respectively, whereas that for untreated PET is 75�0.2˚. In addition, the results<br />

of X-ray photoelectron spectroscopy (XPS) measurements shown in Figure 3a reveal the<br />

existence of a small N1s, Si2s and Si2p peaks at 400, 150 and 100 eV, respectively. These<br />

peaks correspond to 0.6 atom% N and 1.1 atom% Si from the hydrolyzed APTES on the<br />

PET-SiOH surface. In the next step, BMPUS molecules are attached to the PET-SiOH<br />

surface (cf. Figure 1) to serve as initiator centers for the "grafting from" polymerization of<br />

NIPAAm.<br />

211


Subsequent growth of PNIPAAm brushes from the initiator centers at the fiber surface is<br />

verified by the FTIR and XPS spectra presented in Figures 2c and 3b, respectively. The<br />

characteristic secondary amide IR vibrations located at 1650, 1550, 1470, and 3300 cm -1 are<br />

the most pronounced for PET-PNIPAAm microfibers. In addition, the appearance of a<br />

relatively intense N1s peak at 400 eV in Figure 3b indicates an elevated concentration of N,<br />

which is consistent with the presence of PNIPAAm brushes. Quantitation of this spectrum<br />

yields the following atomic concentrations: 76.8�0.4% C, 11.6�0.5% N and 11.6�0.3% O.<br />

These values agree favorably with theoretical concentrations (75.0% C, 12.5% N and 12.5%<br />

O) obtained from the chemical structure of PNIPAAm. The high-resolution C1s spectra<br />

included in the insets of Figures 3a and 3b likewise demonstrate that the PNIPAAm brushes<br />

cover the PET film surface. In Figure 3a, the spectrum displays peaks at 289.0 and 286.6 eV<br />

corresponding to O-C=O and C-O functionalities, respectively. These signature peaks for<br />

PET disappear upon growth of the PNIPAAm brushes, which are responsible for a new peak<br />

at 287.8 eV (N-C=O groups) and a shoulder at 286.1 eV (C-N bonds). 32 Since the XPS<br />

fingerprint for PET is lost upon PNIPAAm brush growth, it can be inferred that the thickness<br />

of the dry brushes is at least the probe depth of XPS (~10 nm). According to ellipsometry<br />

measurements of PET-PNIPAAm films on silicon wafer, the dry thickness of the PNIPAAm<br />

brush after a polymerization time of 30 min is ≈40 nm, which, assuming an average grafting<br />

density of 0.45 chains/nm 2 , corresponds to a molecular weight of ≈48 kDa. 33 Although the<br />

microfibers possess a curved surface, we contend that, on the basis of the brush gyration<br />

212


diameter (≈40 nm) relative to the microfiber diameter (600-1200 nm), the thickness of the<br />

PNIPAAm brush does not differ substantially from that produced on a flat film.<br />

The thermoresponsiveness of the PNIPAAm brushes grown on PET microfibers is first<br />

evaluated with WCA experiments performed successively above and below the Tc of<br />

PNIPAAm, as shown in Figure 4. The WCA of unmodified PET microfibers at 25˚C (Figure<br />

4a) is �125˚, which is higher than that of a flat PET film (75˚) because of the "rough" nature<br />

of the microfiber mat. Despite this increase in surface roughness, the size of the water droplet<br />

on the surface of unmodified PET microfibers does not change during the course of the<br />

measurement, and the measured WCA remains constant. This result also verifies that no<br />

significant evaporation of water takes place during the course of the WCA measurement<br />

because liquid evaporation during WCA measurement may sometimes reduce the apparent<br />

contact angle values due to pinning of the contact line. In Figure 4b, the WCA of the<br />

unmodified PET microfibers at 60˚C is �124˚ and likewise does not change, which suggests<br />

that water evaporation is negligible. Cycling the specimen between these two temperatures in<br />

Figures 4c and 4d yields comparable results, confirming that the PET surface stays<br />

hydrophobic. Measured WCA values of PET-PNIPAAm microfibers, on the other hand,<br />

display significantly different behavior. At 25˚C (Figure 4a), the WCA is also �125˚ when<br />

the water droplet is initially placed on the microfiber surface, but quickly decreases to 0˚ in<br />

just over 40 s as the water is wicked by the hydrophilic PNIPAAm brushes on the surface of<br />

the microfibers. When the temperature is increased beyond Tc of PNIPAAm to 60˚C (Figure<br />

4b), the water droplet is not strongly affected by the microfiber due to the increased<br />

213


hydrophobicity of the PNIPAAm chains, and the WCA is �124˚. Repetition of these<br />

measurements upon thermal cycling in Figures 4c and 4d confirm that the thermorespon-<br />

siveness of PNIPAAm brushes on PET microfibers is reversible with no evidence of<br />

hysteresis.<br />

A second probe of the thermoresponsive nature of PNIPAAm brushes on PET<br />

microfibers employs Au nanoparticles as tracers. Previous studies 29,34 have established that<br />

Au nanoparticles attach to PNIPAAm chains via hydrogen bonding between the citrate<br />

groups present on the nanoparticle surface and the amide groups on PNIPAAm. To discern<br />

the extent to which the PNIPAAm brushes could bind Au nanoparticles, electrospun PET<br />

microfibers have been submerged in a 0.05 mg/ml suspension of Au nanoparticles in<br />

deionized water for 24 h at the same two temperatures examined in Figure 4, i.e., 25 and<br />

60ºC. Images acquired by SEM after drying the fibers reveal that the nanoparticle loading on<br />

the surface of PET-PNIPAAm microfibers is significantly higher at 25ºC (Figure 5a) than at<br />

60ºC (Figure 5b). This difference is attributed to the thermoresponsiveness of the PNIPAAm<br />

chains, which are hydrophilic and swell in water at temperatures below Tc, but become<br />

hydrophobic and collapse in water at temperatures above Tc. As a result of such temperature-<br />

driven swelling or contracting of the brush, the number of NIPAAm units available for<br />

attachment of the Au particles increases or decrease, respectively, which, in turn, governs the<br />

concentration of Au nanoparticles bound to PNIPAAm.<br />

214


Conclusions<br />

In this study, we have demonstrated that the surface of electrospun PET microfibers can<br />

be modified via amidation of the amine group on APTES with the ester group on PET to<br />

permit further chemical modification ultimately resulting in the growth of polymer brushes<br />

by ATRP. Step-by-step examination of the PET surface during the modification sequence,<br />

along with quantitative analysis whenever possible, verifies expectations, and establishes the<br />

sequence as a straightforward and viable route for PET microfiber functionalization. The<br />

thermoresponsive behavior of the PNIPAAm brushes on PET microfibers has been<br />

investigated using both contact angle measurements to determine the nature of the modified<br />

PET surface and Au nanoparticle tracers to determine the extent of brush swelling at<br />

temperatures below and above the lower critical solution temperature of PNIPAAm in water.<br />

Surface functionalization of electrospun PET microfibers using this approach and PNIPAAm<br />

in particular yields mechanically robust and highly porous mats that are temperature-<br />

sensitive, which means that they are suitable candidates for diverse technologies as<br />

responsive filters, scaffolds, delivery vehicles, and sensors.<br />

Acknowledgments: This work was supported by the United Resource Recovery Corporation<br />

and the National Science Foundation through a Graduate Fellowship (K. E. R.).<br />

215


Figures<br />

Figure A1.1. Sequence of surface modification steps employed in this study to functionalize<br />

electrospun PET microfibers with thermoresponsive PNIPAAm brushes. The steps require<br />

deposition and amidation of APTES (a), followed by hydrolysis of the ethoxysilane groups<br />

on APTES to form silanol groups (b), which permit attachment of BMPUS (c) and<br />

subsequent ATRP of NIPAAm to yield PNIPAAm brushes (d). The top and bottom SEM<br />

images display PET and PET-PNIPAAm microfibers, respectively.<br />

216


Figure A1.2. FTIR spectra of (a) as-spun PET, (b) PET-SiOH and (c) PET-PNIPAAm<br />

microfibers. Spectra arranged in the same order in the ex 35 ded views reveal the appearance of<br />

peaks associated with the formation of secondary amide moieties (dotted lines; see text for<br />

assignments).<br />

217


Figure A1.3. XPS spectra of (a) PET-SiOH microfibers and (b) PET-PNIPAAm microfibers.<br />

The survey scans confirm the presence of N upon amidation of PET by APTES in (a) and<br />

PNIPAAm brush formation in (b). The high-resolution insets show the C1s peak (� 285 eV)<br />

before (a) and after (b) PNIPAAm brush growth.<br />

218


Figure A1.4. Cyclic WCA measurements of as-spun PET ( ) and PET-PNIPAAm ( )<br />

microfibers at temperatures (in ºC) below and above the Tc of PNIPAAm: (a) 25, (b) 60, (c)<br />

25, and (d) 60. The error bars correspond to one standard deviation in the data.<br />

219


Figure A1.5. SEM images acquired from PET-PNIPAAm microfibers exposed to aqueous<br />

suspensions of Au nanoparticles at temperatures (labeled) below and above the Tc of<br />

PNIPAAm. The illustrations in the insets (not drawn to scale) portray the conformation of<br />

the PNIPAAm brush at each temperature.<br />

220


References<br />

(1) Taylor, G. Proceedings of the Royal Society of London Series a-Mathematical and Physical<br />

Sciences 1969, 313, 453.<br />

(2) Shin, Y. M.; Hohman, M. M.; Brenner, M. P.et al. Appl. Phys. Lett. 2001, 78, 1149.<br />

(3) Fridrikh, S. V.; Yu, J. H.; Brenner, M. P.et al. Physical Review Letters 2003, 90.<br />

(4) Chae, S. K.; Park, H.; Yoon, J.et al. Advanced Materials 2007, 19, 521; Son, W. K.; Youk, J.<br />

H.; Lee, T. S.et al. Macromolecular Rapid Communications 2004, 25, 1632; Dzenis, Y. Science 2004,<br />

304, 1917.<br />

(5) Sundarrajan, S.; Ramakrishna, S. Journal of Materials Science 2007, 42, 8400.<br />

(6) Yao, C.; Li, X. S.; Neoh, K. G.et al. Appl. Surf. Sci. 2009, 255, 3854.<br />

(7) Lin, J.; Qiu, S. Y.; Lewis, K.et al. Biotechnol. Bioeng. 2003, 83, 168.<br />

(8) Sun, X. Y.; Shankar, R.; Borner, H. G.et al. Advanced Materials 2007, 19, 87; Sun, X. Y.;<br />

Nobles, L. R.; Borner, H. G.et al. Macromolecular Rapid Communications 2008, 29, 1455.<br />

(9) Dann, J. R. Journal of Colloid and Interface Science 1970, 32, 302.<br />

(10) Chen, J. L.; Chu, B.; Hsiao, B. S. Journal of Biomedical Materials Research Part A 2006,<br />

79A, 307.<br />

(11) Muller, K.; Quinn, J. F.; Johnston, A. P. R.et al. Chem. Mater. 2006, 18, 2397.<br />

(12) Ho, C. C.; Chen, W. S.; Shie, T. Y.et al. Langmuir 2008, 24, 5663.<br />

(13) Luong, N. D.; Moon, I. S.; Lee, D. S.et al. Materials Science & Engineering C-Biomimetic<br />

and Supramolecular Systems 2008, 28, 1242; Dong, F. X.; Li, Z. Y.; Huang, H. M.et al. Mater. Lett.<br />

2007, 61, 2556.<br />

(14) Ye, P.; Xu, Z. K.; Wu, J.et al. Biomaterials 2006, 27, 4169.<br />

221


(15) Wang, Z. G.; Wan, L. S.; Xu, Z. K. Soft Matter 2009, 5, 4161.<br />

(16) Winblade, N. D.; Nikolic, I. D.; Hoffman, A. S.et al. Biomacromolecules 2000, 1, 523.<br />

(17) Chronakis, I. S.; Milosevic, B.; Frenot, A.et al. Macromolecules 2006, 39, 357.<br />

(18) Kaur, S.; Ma, Z.; Gopal, R.et al. Langmuir 2007, 23, 13085.<br />

(19) Lee, S.; Obendorf, S. K. Text. Res. J. 2007, 77, 696.<br />

(20) Liao, S.; Murugan, R.; Chan, C. K.et al. Journal of the Mechanical Behavior of Biomedical<br />

Materials 2008, 1, 252.<br />

(21) Zeng, J.; Aigner, A.; Czubayko, F.et al. Biomacromolecules 2005, 6, 1484.<br />

(22) Bhat, R. R.; Tomlinson, M. R.; Wu, T.et al. In Surface- Initiated Polymerization Ii 2006; Vol.<br />

198, p 51.<br />

(23) Kidoaki, S.; Ohya, S.; Nakayama, Y.et al. Langmuir 2001, 17, 2402.<br />

(24) Roux, S.; Demoustier-Champagne, S. Journal of Polymer Science Part A: Polymer Chemistry<br />

2003, 41, 1347; Bech, L.; Elzein, T.; Meylheuc, T.et al. European Polymer Journal 2009, 45, 246.<br />

(25) Farhan, T.; Huck, W. T. S. European Polymer Journal 2004, 40, 1599.<br />

(26) Fadeev, A. Y.; McCarthy, T. J. Langmuir 1998, 14, 5586.<br />

(27) Ellison, M. S.; Fisher, L. D.; Alger, K. W.et al. Journal of Applied Polymer Science 1982, 27,<br />

247; Avny, Y.; Rebenfeld, L. Journal of Applied Polymer Science 1986, 32, 4009.<br />

(28) Bhat, R. R.; Genzer, J. Applied Surface Science 2006, 252, 2549.<br />

(29) Matyjaszewski, K.; Miller, P. J.; Shukla, N.et al. Macromolecules 1999, 32, 8716.<br />

(30) Tomlinson, M. R.; Efimenko, K.; Genzer, J. Macromolecules 2006, 39, 9049; Bhat, R. R.;<br />

Tomlinson, M. R.; Genzer, J. Macromolecular Rapid Communications 2004, 25, 270.<br />

(31) Beamson, G.; Briggs, D. J. High resolution XPS of organic polymers : the Scienta ESCA300<br />

database; Wiley: Chichester England ; New York, 1992.<br />

222


(32) Tomlinson, M. R.; Genzer, J. Langmuir 2005, 21, 11552.<br />

(33) Gupta, S.; Agrawal, M.; Uhlmann, P.et al. Chemistry of Materials, 22, 504<br />

223


APPENDIX II<br />

Generation of Functional PET Microfibers through Surface-Initiated<br />

Abstract<br />

Polymerization<br />

Ali Evren Özçam, Kristen E. Roskov, Richard J. Spontak and Jan Genzer<br />

In this study, we report on a facile and robust method by which poly(ethylene<br />

terephthalate) (PET) surfaces can be chemically modified with functional polymer brushes<br />

while avoiding chemical degradation. The surface of electrospun PET microfibers has been<br />

functionalized by growing poly(dimethylaminoethyl methacrylate) (PDMAEMA) and<br />

poly(2-hydroxyethyl methacrylate) (PHEMA) brushes through a multi-step chemical<br />

sequence that ensures retention of mechanically robust microfibers. Polymer brushes are<br />

grown via "grafting from" atom-transfer radical polymerization after activation of the PET<br />

surface with 3-aminopropyltriethoxysilane. Spectroscopic analyses confirm the expected<br />

reactions at each reaction step, as well as the ultimate growth of brushes on the PET<br />

microfibers. Post-polymerization modification reactions have likewise been conducted to<br />

further functionalize the brushes and impart surface properties of biomedical interest on the<br />

PET microfibers. Antibacterial activity and protein resistance of PET microfibers<br />

functionalized with PDMAEMA and PHEMA brushes, respectively, are demonstrated,<br />

224


thereby making these surface-modified PET microfibers suitable for filtration media, tissue<br />

scaffolds, delivery vehicles, and sensors requiring mechanically robust support media.<br />

Introduction<br />

Electrospinning produces solid polymer fibers with diameters ranging from several tens<br />

of nanometers up to several microns. These nano/microfibers possess a high ratio of surface<br />

area to volume and are typically organized in high-porosity mats that are suitable for use in a<br />

broad range of applications involving (but not limited to) filters, 1 sensors, 2 nanocomposites, 3<br />

tissue engineering scaffolds, 4 drug delivery vehicles, 5 and energy storage media. 6 During<br />

electrospinning, a polymer solution or melt with acceptable viscosity and conductivity levels<br />

is subjected to an electric field acting between a syringe needle and a collector plate. When<br />

electrostatic forces overcome the surface tension of the liquid at the tip of the needle, a<br />

charged polymer solution/melt jet is emitted from the resulting conical structure known as<br />

the Taylor cone. 7 The jet undergoes a whipping process during which any solvent present<br />

evaporates, and the polymer is commonly collected as a dry, randomly oriented fiber mat on<br />

the grounded collector plate. 8 Electrospinning represents an appealing and facile means of<br />

nano/microfiber production due to its relatively straightforward setup and the ability to tune<br />

independent fiber properties on the basis of both solution/melt characteristics and process<br />

parameters.<br />

Although the structural features of electrospun nano/microfibers are advantageous, the<br />

bulk properties of such fibers tend to lack the (multi)functionality that is needed for many<br />

technologies. One way to overcome this problem is to create composite nano/microfibers by<br />

225


incorporating chemically and/or physically different species (molecules or nanoparticles) into<br />

the fibers to enhance, for example, mechanical, 9 electrical, 10 magnetic, 11 or optical 12<br />

properties. Because functional species intended for use on polymer surfaces often exhibit<br />

reduced surface activity when incorporated in a polymer matrix prior to electrospinning, they<br />

may not always locate at the surface where their functionality is desired. 13 For instance, when<br />

antibacterial biocides are added to a polymer prior to electrospinning, their efficacy is greatly<br />

compromised, and they may become unable to attack airborne pathogens. 14 Sun et al. 15 have,<br />

however, established that polarizable peptide-containing copolymers added to a polymer<br />

prior to electrospinning can be brought to the surface of nano/microfibers by the applied<br />

electric field, thereby resulting in fibers that are concurrently electrospun and<br />

biofunctionalized. Alternatively, the surface of electrospun nano/microfibers can be modified<br />

through the covalent bonding of poly(quarternary ammonium), which likewise creates a<br />

permanent antibacterial surface. 16<br />

Because polymer surfaces typically possess low surface energy, they must be pretreated<br />

chemically or physically to obtain an active surface suitable for subsequent<br />

functionalization. 17 Physical methods by which to activate a polymer surface include plasma<br />

treatment, 18 'layer' formation, 19 UV treatment, 14 mineralization, 18 etching, 20 or inclusion of a<br />

reactive composite material. 21 Once chemically-active groups reside on the surface, covalent<br />

bonding, 22 immobilization, 23 and electrostatic interactions 24 can be used to attach reactive<br />

groups to the fiber surface. Modification of only the fiber surface can make commodity and<br />

engineering plastics in particular suitable for applications wherein the fibers interact with<br />

226


their environment, such as molecular filtration, 25 protective textiles, 26 tissue scaffolds, 27 and<br />

drug delivery. 5 A synthetic polymer that shows particular promise in this regard is<br />

poly(ethylene terephthalate) (PET), and electrospun PET microfibers have already been<br />

considered in applications that benefit from the mechanical strength, transparency, and<br />

solvent resistance of PET. 28 As with most organic polymers, however, PET does not possess<br />

good adhesion and wetting properties because of its inherently low surface energy (42<br />

mJ/m 2 ). Application of electrospun PET microfibers as functional materials therefore<br />

necessitates alteration of their surface properties without compromising their bulk<br />

characteristics.<br />

Modification of PET surfaces has been conducted by a variety of chemical methods,<br />

including chemical treatment (e.g., hydrolysis, 29,30,31 reduction, 31,32,33 aminolysis, 30,32,33,34<br />

glycolysis, 31 polyelectrolyte deposition, 35 surface graft polymerization after surface<br />

activation 34,36 ) and physical modification (e.g., plasma, 37,38 ultraviolet/ozone, 38,39 flame, 38<br />

corona treatment, 38,40 electrical discharge, 41 ion beam bombardment, 42 and laser treatment 43 ).<br />

Since most of these surface modification techniques involve, purposefully or inadvertently,<br />

polymer degradation, careful selection of experimental conditions is imperative for the<br />

successful surface modification of PET microfibers without degrading the bulk polymer and<br />

its desirable mechanical properties. Grafting polymer brushes on surfaces represents an<br />

attractive approach by which to modify and control the surface properties of materials.<br />

Surface-initiated graft polymerization has been performed successfully on flat surfaces with a<br />

variety of monomers and polymerization methods. 44 Specifically, atom transfer radical<br />

227


polymerization (ATRP) has been employed 45,46 extensively because (i) its controlled<br />

implementation does not require ultrapurification of the chemicals used, and (ii) it can be<br />

used to polymerize numerous functional monomers, such as N-isopropylacrylamide<br />

(NIPAAm), 47 2-(dimethylamino)ethyl methacrylate (DMAEMA) 48 and 2-hydroxyethyl<br />

methacrylate (HEMA), 49 as well as others.<br />

Graft polymerization on PET surfaces has been reported by Roux and Demoustier-<br />

Champagne 36 and Bech et al. 50 (for styrene), as well as Farhan and Huck 51 (for NIPAAm).<br />

These efforts employ various means of attaching surface initiators for "grafting from"<br />

polymerization. For instance, Roux and Demoustier-Champagne 36 have attached free-radical<br />

polymerization initiators to the surface of PET via electrostatic interactions and covalent<br />

bonding after surface activation by saponification and oxidation. Farhan and Huck 51 and<br />

Bech et al. 50 have alternatively attached ATRP initiators after activating the PET surface by<br />

plasma treatment and aminolysis, respectively. The major drawback of these approaches —<br />

viz., saponification, aminolysis, and plasma treatment — is that they often induce severe<br />

degradation of PET and promote a roughened surface topography. Because of the<br />

nano/micrometer dimensions of electrospun PET fibers, it is paramount that material<br />

degradation and surface roughening must be minimized. In this work, we have elected to<br />

functionalize PET microfiber surfaces by means of 3-aminopropyltriethoxysilane (APTES).<br />

Bùi et al., 32 Fadeev and McCarthy, 52 and Xiang et al. 53 have demonstrated that the primary<br />

amine group in APTES inserts into the PET chain via an amidation reaction with negligible<br />

degradation to bulk PET. In this reaction mechanism the triethoxysilane groups of APTES<br />

228


are exposed at the air interface, and subsequent hydrolysis of the ethoxysilane units yields<br />

silanol groups on the PET surface. These groups are suitable as attachment points for an<br />

ATRP initiator, such as [11-(2-bromo-2-methyl)propionyloxy] undecyltrichlorosilane<br />

(BMPUS). We use this approach to grow PDMAEMA and PHEMA brushes on electrospun<br />

PET microfibers and a battery of analytical techniques to characterize the properties of<br />

electrospun PET microfibers before and after surface polymerization of DMAEMA and<br />

HEMA via ATRP. We also investigate the post-polymerization modification of the<br />

corresponding PDMAEMA and PHEMA brushes via quarternization and fluorination,<br />

respectively, and demonstrate the antibacterial and protein-resistance properties of these<br />

functionalized PET microfiber mats.<br />

Experimental<br />

Materials<br />

Food-grade recycled PET flakes were kindly supplied by United Resource Recovery<br />

Corporation (Spartanburg, SC). Anhydrous toluene, 2-chlorophenol, methanol, iodomethane,<br />

iodopropane, iodobutane, bromoethane, bromopropane, bromobutane, trifluoroacetic<br />

anhydride (TFAA), fibrinogen from human plasma, 1X-PBS buffer (0.137 M NaCl, 0.0027<br />

M KCl, and 0.0119 M phosphates), APTES, DMAEMA, HEMA, Cu(I) bromide (CuBr),<br />

Cu(II) bromide (CuBr2), Cu(I) chloride (CuCl), Cu(II) chloride (CuCl2), bipyridine, and<br />

N,N,N',N',N"-pentamethyldiethylenetriamine (PMDETA) were all purchased from Sigma-<br />

229


Aldrich (St. Louis, MO) and used as-received. Hexafluoroisopropanol (HFIP) was obtained<br />

from Oakwood Products Inc. (Estill, SC).<br />

Brush Growth/Modification<br />

The PET flakes were dissolved in HFIP at different polymer concentrations and<br />

electrospun at 10 kV to generate microfibers possessing different diameters. Thin PET films<br />

measuring 12 and 180 nm thick were likewise spin-coated on silicon wafers from 0.5 and 3<br />

wt% solutions, respectively, in 2-chlorophenol. The latter specimens allowed us to follow<br />

each modification step by measuring film thickness increments associated with the various<br />

chemical modification steps and protein adsorption. Fiber mats and thin films were kept<br />

under vacuum for at least 48 h prior to use to remove entrapped solvent. The initiator for<br />

ATRP was BMPUS, synthesized as described earlier 46 and deposited on the surface of PET<br />

microfibers and films after activation of the surface with APTES. 54 Polymer brushes<br />

composed of either PDMAEMA or PHEMA were subsequently grown from the BMPUS-<br />

decorated PET surfaces by ATRP according to established protocols. 48 For instance, 10.09 g<br />

HEMA was mixed with 6.81 g of methanol, 1.88 g of water and 0.63 g of bipyridine in an<br />

Ar-purged Schlenk flask, and oxygen was removed via three freeze-thaw cycles. After<br />

removal of oxygen, CuCl (0.18 g) and CuCl2 (0.01 g) were added to the solution and 1 more<br />

freeze-thaw cycle was performed. This ATRP solution was transferred to a tightly sealed<br />

Schlenk flask, which was stored in an Ar-purged glove box. Fiber mats and films decorated<br />

with BMPUS were submerged in the ATRP solutions for 6 and 8 h to produce PDMAEMA<br />

and PHEMA brushes, respectively. After removal from the ATRP solution, the samples were<br />

230


insed promptly with methanol and deionized water, and then sonicated in deionized water<br />

for 10 min.<br />

The PDMAEMA brushes grown on PET microfibers and silicon wafers were<br />

quarternized with iodomethane, iodopropane, iodobutane, bromoethane, bromopropane, and<br />

bromobutane in acetonitrile at 60˚C for �20 h. An excess amount of quarternization agents<br />

was added to the glass vial containing PDMAEMA-modified PET microfiber mats and<br />

acetonitrile to yield fully quarternized (q) PDMAEMA brushes. In similar fashion, the<br />

PHEMA brushes were fluorinated with TFAA to discern the effect of fluorinated PHEMA<br />

(fPHEMA) on protein adsorption. In this case, TFAA was used to bind fluorinated moieties<br />

to the hydroxyl terminus of the HEMA pendant group. All reactions were conducted at<br />

ambient temperature in the gas phase, and the samples were washed with copious amounts of<br />

ethanol and water and dried under reduced pressure before protein adsorption experiments. In<br />

both of the post-polymerization modification reactions listed above, bare (i.e., brush-free)<br />

PET microfibers were immersed in the post-polymerization modification reaction mixtures as<br />

controls.<br />

Material Characterization<br />

The thicknesses of the PET films deposited on silicon wafer were measured with<br />

variable-angle spectroscopic ellipsometry (VASE, J.A. Woollam) at an incidence angle of<br />

70º (between the beam and the surface normal) before and after each modification step to<br />

measure the approximate PDMAEMA and PHEMA brush thicknesses on the PET<br />

microfibers. In addition, the thickness of the polymer brushes after quarternization,<br />

231


fluorination, and protein adsorption was also measured with VASE to determine the extent of<br />

these post-polymerization modification steps. The surface chemical composition of modified<br />

microfibers was measured after each modification step by X-ray photoelectron spectroscopy<br />

(XPS) performed on a Kratos Axis Ultra DLD spectrometer at a take-off angle of 90º (under<br />

these conditions the probing depth of XPS is estimated 55 to be ≈9-10 nm). Fourier transform<br />

infrared (FTIR) spectroscopy was utilized to monitor chemical changes that occurred on the<br />

surface of the PET microfibers after modification. Spectra were recorded on a Nicolet 6700<br />

spectrometer after embedding microfiber mats in KBr pellets for analysis in transmission<br />

mode, and resulting data were analyzed by the Omnic Specta software. For each sample,<br />

1024 scans were collected at a resolution of 4 cm -1 . As-spun and surface-modified PET<br />

microfibers were coated with ≈8 nm of Au, and their diameter and surface morphology were<br />

examined by field-emission scanning electron microscopy (SEM) performed on a JEOL<br />

6400F electron microscope operated at 5 kV.<br />

Antibacterial Activity<br />

The PET microfibers decorated with a qPDMAEMA brush were subjected to<br />

antibacterial testing using a modified ASTM standard (E2149-01 Standard Test Method for<br />

Determining the Antimicrobial Activity of Immobilized Antimicrobial Agents under Dynamic<br />

Contact Conditions). Here, E. coli, a model gram-negative bacteria, was grown in a Lauria-<br />

Bertani (LB) medium overnight to yield a bacteria population of 5 x 10 8 according to UV-Vis<br />

spectrophotometry. After serial dilutions, the modified PET microfiber mats (measuring �1<br />

cm 2 in area) were incubated in a suspension containing 3 x 10 5 bacteria in sterile conical<br />

232


tubes at 37�C while being shaken at 300 rpm for 1 h. The resultant suspension was then<br />

diluted with LB medium to a desired concentration and spread on L-agar plates. The L-agar<br />

plates were incubated at 37°C for 18 h. Each surviving cell developed into a distinct bacterial<br />

colony and provided information regarding bacterial activity. The number of viable cells was<br />

measured in terms of colony forming units (CFUs) on each plate.<br />

Protein Resistance<br />

A 0.1 mg/ml solution was prepared at the isoelectric point of fibrinogen (FIB, at pH =<br />

5.5) by dissolving FIB in 1X-PBS buffer solution (0.2% NaN3 was added to the buffer to<br />

prevent bacterial growth). The solution was passed through a 0.2 �m filter, and adsorption<br />

studies of FIB were conducted by incubating substrates in protein solution for 16 h at<br />

ambient temperature. Both fluorinated and unmodified PHEMA brushes on PET microfibers<br />

and films were tested alongside bare PET microfibers and films. After incubation, samples<br />

were washed thoroughly with deionized water, dried under reduced pressure and stored in<br />

glass vials for further characterization. The thickness of the adsorbed FIB layer was<br />

measured by VASE on flat samples (PET films on silicon wafers), and the corresponding<br />

nitrogen surface concentration was measured by XPS to ascertain the amount of adsorbed<br />

FIB.<br />

Results and Discussion<br />

The diameters of electrospun PET microfibers, prepared according to the protocol<br />

provided in the Experimental section and measured by SEM, are 450±100, 800±200 and<br />

1200±300 nm for 6, 8 and 10 wt% solutions, respectively, of PET in HFIP. The surfaces of<br />

233


unmodified PET microfibers appear consistently smooth, as is evident in Figure 1. Functional<br />

PDMAEMA and PHEMA brushes have been grown on PET microfibers in the sequence of<br />

four steps reported earlier 54 for PNIPAAm brushes. Briefly, APTES molecules are attached<br />

to the PET surface via aminolysis between PET and the primary amine of APTES. Next, the<br />

ethoxysilane groups on APTES are hydrolyzed to generate silanol groups for BMPUS<br />

attachment. Finally, PDMAEMA and PHEMA brushes are grown directly from the PET<br />

microfiber surface via ATRP.<br />

As reported previously by Bùi et al. 32 and Fadeev and McCarthy, 52 the primary amine<br />

group in APTES reacts with the ester functionality in PET by forming an amide bond via<br />

aminolysis (cf. Figure 1). Both studies claim that aminolysis of PET with APTES does not<br />

degrade bulk PET as opposed to aminolysis of PET with short alkyl amines, since the latter<br />

can diffuse into a PET fiber and react all the way through, and thus weaken, the fiber. 30,33<br />

The presence of bulky triethoxysilane group on APTES molecules hinders the diffusion of<br />

APTES into PET by increasing the size of the molecule, changing the solubility of the alkyl<br />

amine to which it is attached and creating a protective surface layer (which serves to impede<br />

the diffusion of other APTES molecules). While neither Fadeev and McCarthy 52 nor<br />

Howarter and Youngblood 56 could detect the formation of amide groups on PET due to the<br />

small population of amide groups available on their flat samples, the presence of amide<br />

groups on PET-SiOH microfiber surfaces has been directly confirmed via FTIR analysis by<br />

Ozcam et al. 54 as a result of the enlarged surface area afforded by electrospun PET<br />

microfibers. The appearance of new peaks at 1650 (amide I band), 1550 (amide II band),<br />

234


3300, 1470, and 3300 cm -1 are attributed to the formation of secondary amide groups on the<br />

surface of PET microfibers. Attachment of APTES, followed by hydrolysis of the exposed<br />

triethoxysilane groups in acidic water (pH ≈ 4.5-5.0), yields further reactive silanol groups,<br />

as indicated by both a decrease in water contact angle (WCA) and associated thickness<br />

measurements performed on flat PET films on silicon wafers. For instance, the WCAs of<br />

APTES-modified PET and virgin PET films are 50˚�0.8˚ and 71˚�0.8˚, respectively, after<br />

exposure to acidic water, in contrast to the WCA of native PET (75˚�0.2˚). In addition, XPS<br />

spectra collected from PET-SiOH microfibers confirm the existence of a small nitrogen N1s,<br />

Si2s and Si2p peaks at 400, 150 and 100 eV, respectively, which correspond to 0.6 atom%<br />

nitrogen and 1.1 atom% silicon present on the PET-SiOH surface.<br />

Attachment of BMPUS molecules to the PET-SiOH surface as initiator centers for<br />

"grafting from" polymerization of DMAEMA and HEMA, followed by the ATRP conditions<br />

described in the Experimental section, results in the formation of dry brushes measuring ≈50<br />

and ≈45 nm thick, respectively, as determined by VASE analysis of thin spin-coated PET<br />

films on silicon wafers. Here, we assume that the thicknesses of brushes grown on silicon<br />

wafers is comparable to that produced on the microfibers, since the microfibers are relatively<br />

large and possess negligible curvature on the size scale of the brushes. Examples of PET<br />

microfibers after brush growth are presented in Figure 2 and verify that the chemical<br />

reactions undertaken have no discernible effect on microfiber morphology. Corresponding<br />

FTIR spectra of the polymer brushes, along with spectra acquired from electrospun PET and<br />

PDMAEMA and PHEMA brushes grown directly on silicon wafers, are provided in Figure 3.<br />

235


The spectra of PDMAEMA and PHEMA brushes on silicon wafer are included to point out<br />

the chemical changes that occur on the PET microfibers. Careful comparison of these spectra<br />

confirms that PDMAEMA and PHEMA brushes grew from the surface of PET microfibers.<br />

The appearance of new stretching vibrations located at 2770 and 2820 cm -1 for the<br />

PDMAEMA brush (blue line in Figure 1a), for instance, reflects the C-H bond of the –<br />

N(CH3)2 group of PDMAEMA. Likewise, the increase in peak intensity at �3400 cm -1 for the<br />

PHEMA brush (blue line in Figure 1b) is a consequence of the broad –OH peak originating<br />

from PHEMA.<br />

The chemical compositions of PDMAEMA and PHEMA brushes grown from the surface<br />

of PET microfibers have been assessed by XPS, and resulting values are listed in Table 1.<br />

The theoretical values of these compositions are calculated on the basis of the number of<br />

atoms present on each repeat unit of both polymers and the assumption that the brush<br />

thickness is larger than the probing depth of XPS (�10 nm). Representative XPS spectra of<br />

PET microfibers with grafted PDMAEMA and PHEMA brushes are plotted for comparison<br />

in Figure 4. Bare PET exhibits 2 ionization peaks, one for carbon (at 285 eV) and the other<br />

for oxygen (at 536 eV). Corresponding surface concentrations, computed from the areas<br />

under these curves, are 73.2�0.4 (71.4) and 26.8�0.4 (28.6) atom% for carbon and oxygen,<br />

respectively, which agree favorably with the theoretical values provided in parentheses. On<br />

one hand, growth of a PDMAEMA brush on the surface of PET microfibers is responsible<br />

for the appearance of the nitrogen peak at 400 eV (cf. Figure 4b). Quantitation of such XPS<br />

spectra results in surface concentrations of 73.6�0.5, 7.6�0.1 and 18.8�0.5 atom% for<br />

236


carbon, nitrogen and oxygen, respectively. Introduction of a PHEMA brush, on the other<br />

hand, does not change the number or the position of the peaks recorded in XPS spectra.<br />

Instead, the relative peak areas are affected so that the surface concentrations become<br />

70.3�0.5 and 29.7�0.5 atom% for carbon and oxygen, respectively. These concentration<br />

values measured experimentally for PDMAEMA and PHEMA brushes grown on PET<br />

microfibers are in good quantitative agreement with the theoretical values determined on the<br />

basis of the chemical structures of the individual species (cf. Table 1).<br />

Ellipsometry measurements of flat PET films on silicon wafer, in conjunction with<br />

independent XPS measurements conducted on PET microfiber surfaces, suggest that the<br />

polymer brushes completely cover the PET surfaces, since the dry thicknesses of the brushes<br />

exceed the probing depth of XPS. The characteristic XPS "fingerprint" of PET disappears<br />

from the high-resolution spectra (displayed in the insets of Figure 4) after growing the<br />

PDMAEMA and PHEMA brushes. Introduction of the peak corresponding to the C-N bond<br />

at 286.1 eV serves to broaden the peaks at 285.0 (the C-C bond) and 286.6 eV (the C-O<br />

bond) for the PDMAEMA brush grown on PET microfibers. In the case of the PHEMA<br />

brush, the intensity of the peak located at 286.6 eV is larger than that at 285.0 eV relative to<br />

the XPS spectrum of bare PET. The peak at 290.0 eV, which corresponds to the O-C=O<br />

groups of acrylates, is present for both PDMAEMA and PHEMA brushes grown on the PET<br />

microfibers. 57<br />

Post-polymerization modification reactions have been performed on the PDMAEMA and<br />

PHEMA brushes grafted to the surface of PET microfibers to introduce antibacterial and<br />

237


protein resistance properties, respectively. One of the potential applications of PDMAEMA<br />

after quarternization of the dimethylamino groups on the DMAEMA repeat unit is as an<br />

antibacterial material. Polymer chains quarternized with alkyl halides possess positive<br />

charges and hydrophobic alkyl chains, which induce cation exchange and penetration through<br />

the bacterial cell membrane, respectively. These result in disruption of membrane integrity<br />

and death of bacterial cells. 58 Antibacterial properties of quarternary ammonium compounds<br />

(QACs) have been reported earlier in solution 59 and on solid surfaces. 60,61,62 The latter has an<br />

important advantage over free QACs because they are covalently attached to substrates,<br />

which, in turn, permit repeated use with limited biocidal release to the environment. 63<br />

Quarternization of the PDMAEMA brush grown on the surface of PET microfibers has been<br />

achieved with alkyl bromides differing in length to yield a polycationic brush. Conversely,<br />

the PHEMA brush can be modified to resist protein adsorption, 64 which remains a significant<br />

challenge in biomedical applications involving artificial implants. Adsorption of biomass on<br />

the surface of functional materials degrades the functionality over time. Biomass<br />

accumulation begins with protein adsorption and denaturation on any surface with which<br />

proteins come in contact. Protein adsorption on various surfaces has been studied extensively<br />

over the past several decades, and the incorporation of ethylene glycol and fluorinated units<br />

into polymeric coatings have been found to be among the most effective at reducing the<br />

propensity for protein adsorption.<br />

In the present study, several quarternization agents differing in alkyl length —<br />

iodomethane, iodopropane, iodobutane, bromoethane, bromopropane, and bromobutane —<br />

238


have been used to introduce positive charges into the PDMAEMA brush grown on PET<br />

microfibers and generate a polycationic qPDMAEMA brush with antibacterial properties.<br />

Similarly, TFAA has been used to fluorinate the –OH groups of the PHEMA brush so that<br />

the effect of fPHEMA on the protein resistance of functionalized PET microfiber mats can be<br />

probed. The morphologies of PET microfibers after these post-polymerization modification<br />

reactions are visible in Figure 5 and verify that a microfibrous network of the electrospun<br />

mat is retained. Chemical modification of PDMAEMA and PHEMA brushes grown on<br />

silicon wafers (not on PET film) with quarternization and fluorination agents, respectively,<br />

results in multiple changes in the FTIR spectra. For example, the stretching vibrations<br />

located at 2770 and 2820 cm -1 for the PDMAEMA brush in Figure 6a are attributed to the C-<br />

H bond of the –N(CH3)2 group. These peaks disappear completely after quarternization.<br />

Water absorbed by the more hydrophilic qPDMAEMA brush is responsible for the<br />

appearance of the peak located at �3400 cm -1 . New peaks reflecting the formation of C-CO-<br />

CF bonds (at 1789 cm -1 ) and C-F bonds (at 1224 and 1157 cm -1 ) likewise appear after<br />

conversion of PHEMA to fPHEMA in Figure 6b. Moreover, the –OH groups of PHEMA are<br />

consumed during the fluorination reaction with TFAA, and the peak located at �3400 cm -1<br />

disappears. It is noteworthy that the FTIR spectra collected from PDMAEMA and PHEMA<br />

brushes grown on PET microfibers before and after quarternization (PDMAEMA) and<br />

fluorination (PHEMA) possess the same characteristic peaks in Figure 6, thereby providing<br />

evidence that the brushes grown on the PET microfibers are functionalized.<br />

239


The chemical compositions of brushes grown on PET microfibers and silicon wafers after<br />

quarternization and fluorination reactions have been measured by XPS. Examination of the<br />

corresponding XPS spectra (data not shown) reveals the appearance of new peaks for iodine<br />

or bromine after quarternization of PDMAEMA and fluorine after fluorination of PHEMA.<br />

Quantitation of these spectra yields the surface concentrations listed in Table 2 for<br />

qPDMAEMA (with alkyl bromides) and Table 3 for fPHEMA. Interestingly, the<br />

concentration of bromine in the qPDMAEMA brush grown on PET microfibers is �50%<br />

greater than that in the qPDMAEMA brush on silicon wafer. While this difference was<br />

initially attributed to the adsorption or absorption of alkyl bromides on/in PET, XPS spectra<br />

obtained from bare PET microfibers exposed to the quarternization medium for the same<br />

reaction time reveal no existence of bromine. Therefore, we propose that this difference is a<br />

result of the curved nature of the PET microfibers, which apparently possess a lower steric<br />

hindrance (due to the higher surface area) for the quarternization reaction as compared to a<br />

flat surface. The same trend is also observed for the quarternization reaction of PDMAEMA<br />

brushes with alkyl iodides. In contrast, the extent of fluorination of the PHEMA brush grown<br />

on PET microfibers and silicon wafers is similar, and these values are in agreement with<br />

those reported earlier for TFAA-modified PHEMA brushes. 64 This observation, which differs<br />

from the results obtained for qPDMAEMA brushes, may be due to the smaller size of TFAA<br />

relative to the alkyl halides and the gas-phase reaction of TFAA (wherein TFAA molecules<br />

can diffuse through the brush and completely react with all available –OH groups without<br />

restriction). Arifuzzaman et al. 64 have demonstrated that the surface concentration of TFAA-<br />

240


modified PHEMA brushes on silicon wafer does not change as a function of XPS take-off<br />

angle, which evinces that (i) PHEMA brushes react homogeneously throughout the XPS<br />

probing depth and (ii) the gas-phase reaction of TFAA is quantitative with PHEMA brushes<br />

irrespective of the substrate on which they are grown. We hasten to add that exposure of bare<br />

PET microfibers to TFAA did not alter the surface composition of the bare PET, according to<br />

XPS analysis.<br />

The presence of QACs endows the surface of the PET microfibers with antibacterial<br />

properties due to the presence of cationic groups that disrupt cell membranes and induce<br />

bacterial lysis. 65 In the case of gram negative bacteria such as E. coli, the phosphate groups<br />

of lipopolysaccharide molecules located in the outer bacterial membrane are stabilized by<br />

divalent cations, which would otherwise strongly repel each other, via bridging and<br />

neutralizing. Bacteria lose their natural counterions and their outer membrane is destabilized<br />

upon interacting with QACs due to the electrostatic compensation of these charges with the<br />

cationic charges of the QACs. Thus, the release of counterions from the outer cell wall<br />

initiates the death of the bacteria. 63,65,66 Quarternization of the PDMAEMA brush on PET<br />

microfibers with alkyl bromides differing in methylene length produces string (quenched)<br />

polycationic brushes on the microfiber surface. As reported earlier, 60,61,67 the antibacterial<br />

efficacy of a QAC depends on the extent of quarternization, as well as the length of the alkyl<br />

chain in the quarternization agent. The extent of quarternization dictates the number of<br />

positive charges available to interact with the bacterial membrane, whereas the length of the<br />

alkyl chain affects the antibacterial efficacy by governing the depth of penetration through<br />

241


the cell wall. In general, antibacterial efficiency increases as the length of the alkyl spacer is<br />

increased, but deteriorates after 6 methylene units. 61 As seen in Figure 7a, the presence of a<br />

polycationic qPDMAEMA brush on PET microfibers provides antibacterial properties<br />

against E. coli as the number of CFUs on the agar plates with qPDMAEMA-modified<br />

microfibers is lower than on those with PDMAEMA-modified microfibers. In addition, the<br />

antibacterial efficiency of the qPDMAEMA brush increases substantially with increasing<br />

alkyl length of the quarternization agent from bromoethane to bromobutane, as indicated by<br />

the results provided in Figure 7b.<br />

The resistance of PHEMA and fPHEMA brushes to protein adsorption on flat surfaces<br />

has been recently investigated, and the presence of PHEMA 49 and fPHEMA 64 brushes has<br />

been found to reduce protein adsorption, depending on the graft density and molecular<br />

weight of the brush. In this work, we only examine the protein resistance of a PHEMA brush<br />

grown on PET microfibers before and after fluorination with TFAA. Several different<br />

samples including PHEMA brushes on PET microfibers, fPHEMA brushes on PET<br />

microfibers, PHEMA brushes on silicon wafer, fPHEMA brushes on silicon wafer, bare PET<br />

microfibers and bare PET microfibers exposed to TFAA have all been incubated in FIB<br />

solution for 16 h. Adsorption of FIB on flat substrates has been monitored by measuring the<br />

brush thickness with VASE (on silicon wafers) and the surface nitrogen concentration (due to<br />

FIB) with XPS (on fibers and silicon wafers). Comparing the FIB layer thickness on flat<br />

surfaces reveals that the presence of a PHEMA brush dramatically reduces protein<br />

adsorption. According to the results presented in Figure 8, the thickness of FIB plummets<br />

242


from �4 nm to almost 0 nm after in the presence of a PHEMA brush, and this reduction is<br />

corroborated by XPS data that confirm a corresponding decrease in nitrogen concentration<br />

from 15.1 to 0.6 atom%. The amount of FIB adsorbed on spin-coated PET film is comparable<br />

to that adsorbed on bare silica wafer, but the concentration of adsorbed FIB on bare PET<br />

microfiber (with and without TFAA treatment) is noticeably lower than that on spin-coated<br />

PET film. Introduction of a PHEMA brush on PET microfibers effectively prevents FIB<br />

absorption, but quaternization of the brush does not appear to afford further improvement. In<br />

contrast, fluorination of the PHEMA brush grown on silicon wafer slightly improves protein<br />

resistance, as discerned from both thickness and XPS results.<br />

Conclusions<br />

In this work, we have demonstrated that the surface of electrospun PET microfibers can<br />

be controllably modified via the amidation reaction of the amine group on APTES with the<br />

ester group of PET and subsequent growth of functional polymer brushes (PDMAEMA and<br />

PHEMA) by ATRP. Post-polymerization modification of these brushes has been conducted<br />

by quarternization (PDMAEMA) and fluorination (PHEMA) reactions. None of these brush-<br />

growing or post-functionalization reactions have any discernible deleterious effect on the<br />

morphology of the PET microfibers and, by inference, their robust mechanical properties.<br />

The improved antibacterial efficacy of quarternized PDMAEMA brushes and protein<br />

resistance of (fluorinated) PHEMA brushes grown on PET microfibers are established. These<br />

functional microfiber mats are suitable for use as affinity filters, antibacterial clothing and<br />

responsive sensors. Specifically, we envisage that these modified PET microfiber mats can<br />

243


e employed as multi-use filters for water purification applications, in which case the<br />

stability of such brushes must be ascertained as a function of pH and temperature. In this and<br />

related technologies, the ability of surface-modified microfibers to withstand environmental<br />

stresses is of paramount importance, which is why we have elected to use electrospun PET<br />

microfibers and why we have<br />

chosen a chemical reaction route that does not compromise the mechanical robustness of<br />

PET.<br />

Acknowledgments<br />

This work was supported by the United Resource Recovery Corporation and the National<br />

Science Foundation through a Graduate Fellowship (K. E. R.).<br />

244


Tables<br />

Table A2.1. Compositions of PET microfiber surfaces with grafted PDMAEMA and<br />

PHEMA brushes from XPS analysis..<br />

PET<br />

microfiber<br />

PDMAEMA<br />

brush<br />

PHEMA<br />

brush<br />

Species analyzed<br />

245<br />

Concentration (atom%)<br />

Carbon Nitrogen Oxygen<br />

Theoretical 71.4 � 28.6<br />

Experimental 73.2�0.4 � 26.8�0.4<br />

Theoretical 72.7 9.1 18.2<br />

Experimental 73.6�0.5 7.6�0.1 18.8�0.5<br />

Theoretical 66.7 � 33.3<br />

Experimental 70.3�0.5 � 29.7�0.5


Table A2.2. Compositions of PET microfiber surfaces and silicon wafer modified with<br />

grafted PDMAEMA brushes after quarternization.<br />

Species analyzed<br />

qPDMAEMA on PET microfiber<br />

with bromoethane<br />

qPDMAEMA on silicon wafer<br />

with bromoethane<br />

Bromoethane on PET microfiber<br />

(control)<br />

qPDMAEMA on PET microfiber<br />

with bromopropane<br />

qPDMAEMA on silicon wafer<br />

with bromopropane<br />

Bromopropane on PET microfiber<br />

(control)<br />

qPDMAEMA on PET microfiber<br />

with bromobutane<br />

qPDMAEMA on silicon wafer<br />

with bromobutane<br />

Bromobutane on PET microfiber<br />

(control)<br />

246<br />

Concentration (atom%)<br />

Carbon Nitrogen Oxygen Bromine<br />

73.9 5.3 17.3 3.5<br />

71.2 7.2 19.2 2.4<br />

71.6 � 28.4 �<br />

74.6 5.7 16.5 3.2<br />

72.4 6.7 18.2 2.7<br />

73.7 � 26.3 �<br />

74.0 5.2 17.3 3.5<br />

73.5 6.4 17.8 2.3<br />

72.6 � 27.4 �


Table A2.3. Compositions of PET microfiber surfaces and silicon wafer modified with<br />

grafted PHEMA brushes after fluorination.<br />

Species analyzed<br />

247<br />

Concentration (atom%)<br />

Carbon Oxygen Fluorine<br />

fPHEMA on PET microfiber 54.4 25.0 20.6<br />

fPHEMA on silicon wafer 53.2 24.6 22.2<br />

TFAA-treated PET microfiber<br />

(control)<br />

74.2 25.3 0.5


Figures<br />

Figure A2.1. Synthetic strategy for growing functional polymers on electrospun PET<br />

microfibers. The deposition and subsequent hydrolysis of APTES is followed by attachment<br />

of BMPUS, which provides access to a variety of "grafting from" reactions. The fiber color<br />

and corresponding reactive species are color-matched. The SEM image shows bare<br />

(unmodified) PET microfibers electrospun from HFIP.<br />

248


Figure A2.2. SEM images acquired from electrospun PET microfibers with (a) PDMAEMA<br />

and (b) PHEMA brushes. The solution concentrations used to generate the PET microfibers<br />

was 8 wt%.<br />

249


Figure A2.3. FTIR spectra collected from systems with (a) PDMAEMA and (b) PHEMA<br />

brushes. Spectra correspond to bare electrospun PET microfibers (dotted black lines),<br />

brushes grown on silicon wafers (solid black lines) and brushes grown on the PET<br />

microfibers (solid blue lines).<br />

250


Figure A2.4. XPS spectra collected from (a) electrospun PET microfibers, as well as PET<br />

microfibers functionalized with (b) PDMAEMA and (c) PHEMA brushes. A high-resolution<br />

carbon-edge spectrum is included in the inset of each panel.<br />

251


FigureA2.5. SEM images acquired from electrospun PET microfibers with postfunctionalized<br />

(a) qPDMAEMA (using bromobutane) and (b) fPHEMA brushes. The<br />

solution concentrations used to generate the PET microfibers were 8 wt%.<br />

252


Figure A2.6. FTIR spectra collected from systems with (a) PDMAEMA and (b) PHEMA<br />

brushes. Spectra correspond to original and post-functionalized brushes grown on silicon<br />

wafers (black and red, respectively), as well as original and post-functionalized brushes<br />

grown on PET microfibers (blue and green, respectively).<br />

253


Figure A2.7. In (a), photographs of E. coli colonies on L-agar plates containing bare PET<br />

microfibers, as well as PET microfibers modified with a PDMAEMA brush and postquaternized<br />

with bromoethane, bromopropane and bromobutane (labeled) after an incubation<br />

period of 18 h at 37°C. The dependence of the number of colony forming units (CFUs) on<br />

the length of the alkyl bromide used is evident in (b). For reference, the CFU corresponding<br />

to the bare PET microfibers is indicated by the red line.<br />

254


Figure A2.8. Surface nitrogen concentration (left ordinate, red bars) and fibrinogen thickness<br />

(right ordinate, blue squares) for various systems containing silicon wafer, electrospun PET<br />

microfibers, PHEMA brushes and post-functionalized fPHEMA brushes.<br />

255


References<br />

[1] R. Gopal, S. Kaur, C. Y. Feng, C. Chan, S. Ramakrishna, S. Tabe, T. Matsuura, J. Membr.<br />

Sci. 2007, 289, 210.<br />

[2] S. K. Chae, H. Park, J. Yoon, C. H. Lee, D. J. Ahn, J.-M. Kim, Adv. Mater. 2007, 19, 521.<br />

[3] R. H. Baughman, A. A. Zakhidov, W. A. de Heer, Science 2002, 297, 787.<br />

[4] B. M. Min, G. Lee, S. H. Kim, Y. S. Nam, T. S. Lee, W. H. Park, Biomaterials 2004, 25,<br />

1289.<br />

[5] J. Zeng, A. Aigner, F. Czubayko, T. Kissel, J. H. Wendorff, A. Greiner, Biomacromol. 2005,<br />

6, 1484.<br />

[6] (a) Y. Dzenis, Science 2004, 304, 1917; (b) L. Chen, L. Bromberg, J. A. Lee, H. Zhang, H.<br />

Schreuder-Gibson, P. Gibson, J. Walker, P. T. Hammond, T. A. Hatton, G. C. Rutledge, Chem.<br />

Mater., 22, 1429.<br />

[7] G. Taylor, Proc. Royal Soc. A (London) 1969, 313, 453.<br />

[8] Y. M. Shin, M. M. Hohman, M. P. Brenner, G. C. Rutledge, Appl. Phys. Lett. 2001, 78,<br />

1149.<br />

[9] F. Ko, Y. Gogotsi, A. Ali, N. Naguib, H. H. Ye, G. L. Yang, C. Li, P. Willis, Adv. Mater.<br />

2003, 15, 1161.<br />

[10] E. J. Ra, K. H. An, K. K. Kim, S. Y. Jeong, Y. H. Lee, Chem. Phys. Lett. 2005, 413, 188.<br />

[11] A. Wang, H. Singh, T. A. Hatton, G. C. Rutledge, Polymer 2004, 45, 5505.<br />

[12] K. E. Roskov, K. A. Kozek, W.-C. Wu, R. K. Chhetri, A. L. Oldenburg, R. J. Spontak, J. B.<br />

Tracy, Langmuir 2011 (in press).<br />

[13] K. M. Sawicka, P. Gouma, J. Nanoparticle Res. 2006, 8, 769.<br />

256


[14] C. Yao, X. S. Li, K. G. Neoh, Z. L. Shi, E. T. Kang, Appl. Surf. Sci. 2009, 255, 3854.<br />

[15] (a) X.-Y. Sun, R. Shankar, H. G. Boerner, T. K. Ghosh, R. J. Spontak, Adv. Mater. 2007, 19,<br />

87; (b) X.-Y. Sun, L. R. Nobles, H. G. Boerner, R. J. Spontak, Macromol. Rapid Commun. 2008, 29,<br />

1455.<br />

[16] J. Lin, S. Y. Qiu, K. Lewis, A. M. Klibanov, Biotechnol. Bioeng. 2003, 83, 168.<br />

[17] J. P. Deng, L. F. Wang, L. Y. Liu, W. T. Yang, Prog. Polym. Sci. 2009, 34, 156.<br />

[18] J. L. Chen, B. Chu, B. S. Hsiao, J. Biomed. Mater. Res. A 2006, 79A, 307.<br />

[19] K. Muller, J. F. Quinn, A. P. R. Johnston, M. Becker, A. Greiner, F. Caruso, Chem. Mater.<br />

2006, 18, 2397.<br />

[20] C. C. Ho, W. S. Chen, T. Y. Shie, J. N. Lin, C. Kuo, Langmuir 2008, 24, 5663.<br />

[21] (a) N. D. Luong, I. S. Moon, D. S. Lee, Y. K. Lee, J. D. Nam, Mater. Sci. Eng. C 2008, 28,<br />

1242; (b) F. X. Dong, Z. Y. Li, H. M. Huang, F. Yang, W. Zheng, C. Wang, Mater. Lett. 2007, 61,<br />

2556.<br />

[22] P. Ye, Z. K. Xu, J. Wu, C. Innocent, P. Seta, Biomaterials 2006, 27, 4169.<br />

[23] Z. G. Wang, L. S. Wan, Z. K. Xu, Soft Matter 2009, 5, 4161.<br />

[24] N. D. Winblade, I. D. Nikolic, A. S. Hoffman, J. A. Hubbell, Biomacromol. 2000, 1, 523.<br />

[25] S. Kaur, Z. Ma, R. Gopal, G. Singh, S. Ramakrishna, T. Matsuura, Langmuir 2007, 23,<br />

13085.<br />

[26] S. Lee, S. K. Obendorf, Textile Res. J. 2007, 77, 696.<br />

[27] S. Liao, R. Murugan, C. K. Chan, S. Ramakrishna, J. Mech. Behavior Biomed. Mater. 2008,<br />

1, 252.<br />

[28] I. S. Chronakis, B. Milosevic, A. Frenot, L. Ye, Macromolecules 2006, 39, 357.<br />

257


[29] (a) J. Dave, R. Kumar, H. C. Srivastava, J. Appl. Polym. Sci. 1987, 33, 455; (b) M. S. Ellison,<br />

L. D. Fisher, K. W. Alger, S. H. Zeronian, J. Appl. Polym. Sci. 1982, 27, 247; (c) E. M. Saunders, S.<br />

H. Zeronian, J. Appl. Polym. Sci. 1982, 27, 4477; (d) W. Chen, T. J. McCarthy, Macromolecules<br />

1998, 31, 3648.<br />

[30] (a) L. N. Bui, M. Thompson, N. B. McKeown, A. D. Romaschin, P. G. Kalman, The Analyst<br />

1993, 118, 463; (b) Y. Avny, L. Rebenfeld, J. Appl. Polym. Sci. 1986, 32, 4009.<br />

[31] R. Fukai, P. H. R. Dakwa, W. Chen, J. Polym. Sci. A: Polym. Chem. 2004, 42, 5389.<br />

[32] L. Dauginet, A. S. Duwez, R. Legras, S. Demoustier-Champagne, Langmuir 2001, 17, 3952.<br />

[33] S. Roux, S. Demoustier-Champagne, J. Polym. Sci. A: Polym. Chem. 2003, 41, 1347.<br />

[34] (a) Y. L. Hsieh, E. Y. Chen, Ind. Eng. Chem. Prod. Res. Dev. 1985, 24, 246; (b) M. Strobel,<br />

M. J. Walzak, J. M. Hill, A. Lin, E. Karbashewski, C. S. Lyons, J. Adhes. Sci. Tech. 1995, 9, 365.<br />

[35] (a) J. M. Hill, E. Karbashewski, A. Lin, M. Strobel, M. J. Walzak, J. Adhes. Sci. Tech. 1995,<br />

9, 1575; (b) M. J. Walzak, S. Flynn, R. Foerch, J. M. Hill, E. Karbashewski, A. Lin, M. Strobel, J.<br />

Adhes. Sci. Tech. 1995, 9, 1229; (c) D. O. H. Teare, C. Ton-That, R. H. Bradley, Surf. Interface Anal.<br />

2000, 29, 276; (d) C. Ton-That, D. O. H. Teare, P. A. Campbell, R. H. Bradley, Surface Science<br />

1999, 435, 278; (e) A. E. Ozcam, K. Efimenko, C. Jaye, R. J. Spontak, D. A. Fischer, J. Genzer, J.<br />

Electr. Spectr. Related Phen. 2009, 172, 95.<br />

[36] (a) M. Strobel, C. S. Lyons, J. M. Strobel, R. S. Kapaun, J. Adhes. Sci. Tech. 1992, 6, 429;<br />

(b) J. M. Pochan, L. J. Gerenser, J. F. Elman, Polymer 1986, 27, 1058.<br />

[37] D. Briggs, D. G. Rance, C. R. Kendall, A. R. Blythe, Polymer 1980, 21, 895.<br />

[38] P. Bertrand, Y. Depuydt, J. M. Beuken, P. Lutgen, G. Feyder, Nucl. Instr. Meth. Phys. Res. B<br />

1987, 19-2, 887.<br />

258


[39] E. Arenholz, J. Heitz, M. Wagner, D. Bauerle, H. Hibst, A. Hagemeyer, Appl. Surf. Sci.<br />

1993, 69, 16.<br />

[40] R. Barbey, L. Lavanant, D. Paripovic, N. Schuwer, C. Sugnaux, S. Tugulu, H. A. Klok,<br />

Chem. Rev. 2009, 109, 5437.<br />

[41] (a) K. Matyjaszewski, H. Dong, W. Jakubowski, J. Pietrasik, A. Kusumo, Langmuir 2007, 23,<br />

4528; (b) K. Matyjaszewski, P. J. Miller, N. Shukla, B. Immaraporn, A. Gelman, B. B. Luokala, T.<br />

M. Siclovan, G. Kickelbick, T. Vallant, H. Hoffmann, T. Pakula, Macromolecules 1999, 32, 8716.<br />

[42] Y. K. Jhon, R. R. Bhat, C. Jeong, O. J. Rojas, I. Szleifer, J. Genzer, Macromol. Rapid<br />

Commun. 2006, 27, 697.<br />

[43] R. R. Bhat, J. Genzer, Appl. Surf. Sci. 2006, 252, 2549.<br />

[44] R. R. Bhat, B. N. Chaney, J. Rowley, A. Liebmann-Vinson, J. Genzer, Adv. Mater. 2005, 17,<br />

2802.<br />

[45] L. Bech, T. Elzein, T. Meylheuc, A. Ponche, M. Brogly, B. Lepoittevin, P. Roger, Eur.<br />

Polym. J. 2009, 45, 246.<br />

[46] T. Farhan, W. T. S. Huck, Eur. Polym. J. 2004, 40, 1599.<br />

[47] A. Y. Fadeev, T. J. McCarthy, Langmuir 1998, 14, 5586.<br />

[48] J. H. Xiang, P. X. Zhu, Y. Masuda, K. Koumoto, Langmuir 2004, 20, 3278.<br />

[49] A. E. Ozcam, K. E. Roskov, J. Genzer, R. J. Spontak (submitted).<br />

[50] J. C. Vickerman, I. Gilmore, Surface Analysis: The Principal Techniques, 2nd ed., John<br />

Wiley & Sons, Ltd., Chichester, 2009.<br />

[51] J. A. Howarter, J. P. Youngblood, Macromolecules 2007, 40, 1128.<br />

[52] G. Beamson, D. J. Briggs, High Resolution XPS of Organic Polymers: The Scienta ESCA300<br />

Database, Wiley, Chichester, 1992.<br />

259


[53] K. Lewis, A. M. Klibanov, Trends Biotechnol. 2005, 23, 343.<br />

[54] S. Lenoir, C. Pagnoulle, C. Detrembleur, M. Galleni, R. Jerome, J. Polym. Sci. A: Polym.<br />

Chem. 2006, 44, 1214.<br />

[55] (a) J. Huang, R. R. Koepsel, H. Murata, W. Wu, S. B. Lee, T. Kowalewski, A. J. Russell, K.<br />

Matyjaszewski, Langmuir 2008, 24, 6785; (b) J. Lin, S. Y. Qiu, K. Lewis, A. M. Klibanov,<br />

Biotechnol. Prog. 2002, 18, 1082; (c) H. Murata, R. R. Koepsel, K. Matyjaszewski, A. J. Russell,<br />

Biomaterials 2007, 28, 4870; (d) J. Y. Huang, H. Murata, R. R. Koepsel, A. J. Russell, K.<br />

Matyjaszewski, Biomacromol. 2007, 8, 1396.<br />

[56] R. Kugler, O. Bouloussa, F. Rondelez, Microbiology-Sgm 2005, 151, 1341.<br />

[57] S. Arifuzzaman, A. E. Ozcam, K. Efimenko, D. A. Fischer, J. Genzer, Biointerphases 2009,<br />

4, FA33.<br />

[58] P. Gilbert, L. E. Moore, J. Appl. Microbiol. 2005, 99, 703.<br />

[59] Y. Endo, T. Tani, M. Kodama, Appl. Environ. Microbiol. 1987, 53, 2050.<br />

[60] J. A. Lichter, K. J. Van Vliet, M. F. Rubner, Macromolecules 2009, 42, 8573.<br />

260


APPENDIX III<br />

Modification of Melt-spun Isotactic Polypropylene and Poly(lactic acid)<br />

Abstract<br />

Bicomponent Filaments with a Premade Block Copolymer*<br />

Sara A. Arvidson, Kristen E. Roskov, Jaimin J. Patel, Richard J. Spontak,<br />

Russell E. Gorga and Saad A. Khan<br />

While numerous studies have investigated the effect of adding a block copolymer as a<br />

macromolecular surfactant to immiscible polymer blends, no such efforts have sought to alter<br />

the properties of melt-spun bicomponent core-sheath filaments with a nonreactive<br />

compatibilizing agent. In this study, we examine the effect of adding poly[styrene-b-<br />

(ethylene-co-butylene)-b-styrene] (SEBS) triblock copolymer to core-sheath filaments<br />

consisting of isotactic polypropylene (iPP) and poly(lactic acid) (PLA). Incorporation of the<br />

copolymer into blends of iPP/PLA is observed to reduce the size scale of phase separation.<br />

Interfacial slip between molten iPP and PLA layers is evaluated by rheology under steady-<br />

shear conditions. Addition of SEBS to the PLA sheath during filament formation reduces the<br />

tendency of PLA sheaths to crack prior to iPP core failure during tensile testing. In reversed<br />

filament configurations, the copolymer does not hinder the development of molecular<br />

orientation, related to fiber strength, during fiber spinning. Electron microscopy reveals that<br />

the copolymer molecules form unique, highly nonequilibrium morphologies under the<br />

spinning conditions employed here.<br />

261


Introduction<br />

Bicomponent filament spinning involves the co-extrusion of two polymers (or polymer<br />

blends) from the same spinneret into a single filament that consists of both starting materials.<br />

This type of spinning can be configured in a variety of cross-sectional geometries such as<br />

core-sheath, side-by-side, segmented pie, islands in the sea, or trilobal. 1 In this study, we only<br />

consider further the core-sheath arrangement. The benefits of co-spinning two polymers<br />

include a reduction in the cost associated with a single-step process, a net increase in the<br />

performance of fibers and webs as derived from the desirable characteristics of the<br />

constituent polymers, and suppression of an unfavorable rheological behavior (e.g., spinning<br />

a polymer behaving as a Newtonian liquid with a small fraction of a desirable polymer<br />

exhibiting Maxwell properties). 2 Moreover, since the viscoelasticity of the sheath polymer<br />

dictates the mechanics of melt flow during coaxial spinning even if its flow rate and viscosity<br />

are lower than those of the core polymer, core-sheath bicomponent fibers may be<br />

significantly thinner than would otherwise be achieved by spinning the polymers<br />

individually. 2 Lastly, while the bicomponent spinning of two polymers permits retention of<br />

component properties, comparable spinning of pre-mixed polymer blends generally results in<br />

properties that lie intermediate between those of the individual components, thus<br />

compromising the net properties of the fibers. 3<br />

Depending on the thermodynamic compatibility of the polymers involved, bicomponent<br />

spinning can result in filaments or fibers (these words are used interchangeably throughout<br />

text) that fail by splitting when subjected to an external mechanical force. 4 While splittable<br />

262


fibers may be desired in the manufacture of synthetic suede and leather, technical wipes, and<br />

some filtration applications, 5 good adhesion between the individual species comprising<br />

bicomponent fibers is required for maintaining mechanical integrity, developing sutures, or<br />

improving chemical or flammability resistance. 6,7 At the interface separating immiscible<br />

polymers, a relatively low population of entanglements can lead to fiber delamination at<br />

temperatures below the melting temperatures (Tms) of both components. Above the Tm of<br />

each component, subjecting immiscible polymers to external forces may result in “slip” of<br />

one molten polymer across the other. Evaluating the presence of slip and, by inference, the<br />

adequacy of interfacial chain entanglements in the melt, remains a nontrivial task. Zhao and<br />

Macosko 8 have deduced that slip of multilayered samples occurs when a drop in viscosity<br />

accompanies an increase in the number of layers and, hence, interfacial contact area. Jiang 9<br />

has similarly interpreted that a viscosity below the theoretical viscosity discerned from the<br />

reciprocal rule of mixtures (R-ROM) for layered high-density polyethylene/polystyrene<br />

(HDPE/PS) is indicative of slip. Slip has also been proposed as the cause for the viscosity<br />

discontinuity encountered while shearing layered HDPE/PS filled with tracer particles and<br />

observing the layers in situ with confocal microscopy. 10 Park et al. 11 have likewise<br />

investigated multilayer slip by using a sliding plate rheometer equipped with a camera.<br />

Block copolymers, composed of two or more long, covalently-linked sequences of<br />

chemically-dissimilar repeat units, may be used to lower interfacial tension along polymer-<br />

polymer interfaces and thus improve adhesion by promoting chain entanglements. The<br />

apparent results of such compatibilization are a net reduction in slip during extrusion and<br />

263


delamination in formed fibers. 12 Generally speaking, compatibilization is considered to be<br />

effective if an added block copolymer reduces the size scale of phase domains and/or<br />

enhances the mechanical properties of a blend. 3,13 The stabilizing efficacy and spatial<br />

segregation of premade block copolymer molecules along polymer-polymer interfaces has<br />

been previously addressed by Wei et al. 22 and more recently by Gozen et al. 23 Alternatively,<br />

reactive compatibilization of polymer blends can be induced through the use of species that<br />

react along the polymer-polymer interface to form block copolymers in situ. 24 In this spirit,<br />

blends of isotactic polypropylene (iPP) and poly(lactic acid) (PLA) have been modified with<br />

a maleic anhydride (MA)-grafted copolymer, which has been shown 25 to increase the impact<br />

strength of the resultant alloy. Similarly, PP-g-MA has also been reported 26 to reactively<br />

crosslink core-sheath and side-by-side bicomponent fibers composed of nylon-6 and iPP. The<br />

concept of compatibilizing polymer-polymer interfaces that are not blended but rather<br />

contacted, such as those involved in core-sheath fibers or laminates, with premade block<br />

copolymers has largely been ignored. To the best of our knowledge, no prior studies<br />

investigating the core-sheath compatibilization of bicomponent fibers by the addition of a<br />

block copolymer to one component have been reported.<br />

During large-scale mechanical deformation, melt-spun bicomponent fibers composed of<br />

iPP and PLA are observed to possess interfacial voids that extend up to millimeters in length<br />

along the fiber axis. These voids are attributed to inherently poor interfacial adhesion<br />

between iPP and PLA. 4 In this work, we endeavor to compatibilize and thus improve the<br />

mechanical properties of these two non-blended polymers by incorporating a triblock<br />

264


copolymer during melt spinning. In addition to filament property assessment, the<br />

morphology of the block copolymer in filaments is compared to that formed during melt<br />

mixing without extrusion to elucidate the effect of high-shear spinning on block copolymer<br />

structuration. Electron microscopy reveals that unique, nonequilibrium copolymer<br />

morphologies are generated during bicomponent filament spinning. We also discuss the<br />

application of steady-shear rheological methods to evaluate slip at layered iPP-PLA<br />

interfaces and the effect of adding the block copolymer.<br />

Experimental<br />

Materials and Specimen Preparation<br />

The iPP and PLA were provided by Sunoco Chemicals (Pittsburgh, PA; CP360H) and<br />

NatureWorks (Minnetonka, MN; 6202D), respectively. A premixed compound containing a<br />

poly[styrene-b-(ethylene-co-butylene)-b-styrene] (SEBS) triblock copolymer with 18.6 wt%<br />

styrene (according to the manufacturer) and an overall molecular weight of 67 kDa<br />

(according to independent size exclusion chromatography) was supplied by Kraton Polymers<br />

(Houston, TX). While the identity of the compounding material is proprietary, it is midblock-<br />

selective, which indicates that it is more aliphatic than aromatic in nature. We do not,<br />

however, discount the possibility that the compounding species (hereafter referred to as the<br />

"midblock extender") may be partially unsaturated, unlike the EB midblock of the<br />

copolymer. Reagent-grade dichloromethane (DCM) was purchased from Mallinckrodt<br />

Chemicals (Phillipsburg, NJ) and used as-received.<br />

265


Single and bicomponent filaments were melt-spun at various aspirator pressures on the<br />

Partners’ Pilot Spunbond line located in the Nonwovens Cooperative Research Center at<br />

North Carolina State University. All filaments examined here were spun with a total mass<br />

throughput of 0.4 g/hole-min at a fiber composition of 50/50 (w/w) core/sheath. In select<br />

cases, 5 wt% of the PLA was replaced with the SEBS copolymer. In these instances, the PLA<br />

and copolymer were melt-compounded and co-extruded. Confluence of the molten PLA (or<br />

PLA + SEBS) and iPP occurred in the spin pack. Below the spin pack, bicomponent<br />

filaments were directed through the quench zone to an attenuation zone, where the aspirator<br />

pressure controlled the air velocity around the fibers and effectively the fiber spinning<br />

velocity. Non-bonded fibers were collected immediately following extrusion so that the as-<br />

spun fiber morphology could be examined. Fiber diameters were used to calculate the<br />

"spinning velocity" (V) at the point where the fibers solidified according to<br />

V � Q<br />

�A c<br />

where Q is the mass flow rate of polymer per spinneret hole, ρ is the fiber mass density, and<br />

Ac is the cross-sectional �� area of the fiber. For comparison with the bicomponent fibers,<br />

corresponding polymer blends were prepared using a Haake-Buchler HBI System 90 twin-<br />

screw melt mixer operated at 185°C. Binary iPP/PLA blends were immersed in DCM for<br />

~100 h under constant agitation at ambient temperature to selectively dissolve the PLA for<br />

morphological evaluation.<br />

266<br />

(1)


Specimen Characterization<br />

X-ray diffraction (XRD) studies were conducted on a Bruker D-5000 diffractometer<br />

(Madison, WI) equipped with a Highstar area detector and using CuKa radiation (λ = 0.1542<br />

nm) at 40 kV and 30 mA. Resultant 2-dimensional XRD patterns were normalized with<br />

respect to an empty sample holder and analyzed with the Bruker General Area Detector<br />

Diffraction System (GADDS) software. Mechanical testing of fibers was conducted at<br />

ambient temperature on an Instron Model 5544 extensiometer (Norwood, MA) fitted with a 5<br />

N load cell. Single filaments with a gauge length of 28.6 mm were strained at a constant<br />

crosshead speed of 25.4 mm/min. Data were analyzed with the Bluehill v.2 software<br />

package, and a constant volume cylinder was assumed to calculate true stress. Representative<br />

stress-strain curves were obtained for specimens prepared at different fiber configurations<br />

and pressures after at least 10 trials. Optical microscopy of bicomponent fibers was<br />

performed on a Mach-Zehnder type interference microscope by Aus<br />

Jena (Jena, Germany) with polarized light (λ= 546 nm), and digital images were collected on<br />

a CCD camera for birefringence analysis. Both scanning and transmission electron<br />

microscopies (SEM and TEM, respectively) were employed to explore the morphologies of<br />

the fibers and blends prepared here. After using DCM to dissolve the PLA from melt-<br />

processed iPP/PLA blends, the remaining iPP matrix was sputter-coated with ~10 nm of Au<br />

and subsequently analyzed by SEM performed in an environmental FEI XL-30 microscope<br />

operated under high vacuum at 5 kV. The average size and standard deviation of the pores<br />

267


introduced by extracting PLA were discerned by measuring the diameter of 100 pores using<br />

the ImageJ software suite.<br />

For complementary TEM examination of the block copolymer morphology, fibers were<br />

conformally sputter-coated with 30 nm of Au (as a barrier layer to avoid fiber<br />

contamination), embedded in epoxy and microtomed at ambient temperature on a Leica<br />

UltraCut 7 with a diamond knife. Resultant sections were stained for 7 min with the vapor of<br />

0.5% RuO4(aq), which is a selective stain for the phenyl groups on the S blocks of the SEBS<br />

copolymer. Cross-sectional TEM images were acquired on a field-emission Hitachi HF2000<br />

microscope operated at an accelerating voltage of 200 kV. Average microdomain sizes and<br />

their corresponding standard deviations were determined by measuring 50-100 features of<br />

interest, unless otherwise noted, using ImageJ. The zero-shear viscosities of the iPP and PLA<br />

homopolymers with and without copolymer, as well as multilayered samples thereof, were<br />

measured at 185°C under nitrogen on a TA Instruments AR-G2 rheometer equipped with<br />

parallel plates measuring 25 or 40 mm in diameter. Multilayered specimens consisted of 1, 2,<br />

4, or 8 alternating layers of iPP and PLA. Dynamic rheology was likewise performed on the<br />

same instrument operated at 185°C with 25 mm plates and a 1 mm gap. Differential scanning<br />

calorimetry (DSC) was conducted on a TA Instruments Q2000 model calorimeter calibrated<br />

to an indium standard. Scans were carried out at heating rates of 10°C/min under 50 mL/min<br />

N2 purge with samples of approximately 10 mg in standard aluminum pans.<br />

268


Results and Discussion<br />

Bicomponent iPP/PLA Fibers<br />

Lipscomb 27 predicts that, for situations wherein the core polymer undergoes a greater<br />

increase in viscosity during cooling than the sheath polymer, the core polymer bears more of<br />

the spinline tension, which we hasten to add can serve to enhance crystallization and<br />

molecular orientation. Similarly, Kikutani 6 reports that the solidification temperature and<br />

viscosity disparities in co-spun polymers constitute the main factors influencing the "mutual<br />

interactions" of the constituent polymers. This conclusion is interpreted to relate to the ability<br />

of one polymer to direct or influence the crystallization and molecular orientation of the other<br />

polymer. On one hand, we find that the crystallization of PLA is not strongly influenced by<br />

its location (i.e., core or sheath) in bicomponent fibers, as seen in Figure A3.1. The iPP<br />

crystal morphology, on the other hand, is sensitively affected by the fiber configuration. Our<br />

observation that iPP tends to crystallize in the sheath but not in the core, coupled with the<br />

results of Lipscomb 27 and our previous work 28 on the quiescent and stress-induced<br />

crystallization of iPP, suggests that iPP crystallization is due to exposure to the quench air<br />

while in the sheath and is not stress-induced while in the core. The presence of the iPP<br />

mesomorphic phase for most spinning velocities indicates an absence of high tension on the<br />

iPP core when co-spun with a PLA sheath (this may reflect the comparable melting<br />

temperatures and zero-shear viscosities of iPP and PLA, as listed in Table A3.1). In marked<br />

contrast, Kikutani et al. 10 have demonstrated that co-spinning iPP with poly(ethylene<br />

terephthalate) (PET) into bicomponent iPPsheath/PETcore fibers results in sheaths that do not<br />

269


solidify in the draw zone and fibers that could not be drawn as finely as either constituent<br />

polymer. Since the difference in melting points between iPP and PET is significantly greater<br />

(~100°C) than that between iPP and PLA, the reason why this problem is not encountered<br />

here is because the melt viscosities and thermal transitions of iPP and PLA are sufficiently<br />

similar so that neither polymer significantly influences the rheological properties or<br />

solidification behavior of the other, compared to each polymer when spun individually.<br />

According to the results displayed in Figure A3.1, an increase in aspirator pressure is<br />

generally accompanied by a reduction (that is significant from 0 to 10 psi) in fiber diameter<br />

for both single-component and bicomponent fibers. At a given aspirator pressure, the<br />

diameters of bicomponent fibers are often marginally lower than the fiber diameters of either<br />

polymer spun individually. This observation, along with the inherent thermodynamic<br />

incompatibility between iPP and PLA (discussed further in the next section), implies that the<br />

iPP/PLA interface may slip at high spin speeds, which occur at high shear rates. We return to<br />

address the occurrence of slip between molten iPP and PLA, as determined by rheology,<br />

later. Addition of the SEBS copolymer to PLA in bicomponent fibers is expected to influence<br />

the iPP/PLA interface and, by inference, the fiber morphology. Figure A3.2 shows the fiber<br />

diameters of bicomponent fibers co-spun with and without copolymer. As is evident in<br />

Figures A3.1 and A3.2, an increase in aspirator pressure generally promotes a decrease in<br />

fiber diameter. Compared to iPP spun alone, finer fibers result from co-spinning an iPP<br />

sheath around a PLA core (Figure A3.2a), but the addition of copolymer appears to have a<br />

non-systematic effect on fiber diameter. With PLA as the sheath (Figure A3.2b),<br />

270


icomponent fibers are slightly smaller in diameter at moderate spin speeds, and the addition<br />

of copolymer has little statistical effect on fiber diameter over the conditions examined.<br />

Compatibilized iPP/PLA Blends<br />

The interfacial energy between iPP and PLA is related to the surface energy of each polymer<br />

according to Antonoff's rule:<br />

� iPP�PLA � � iPP � � PLA (2)<br />

From Eq. 2, the interfacial energy between iPP and PLA is estimated to be about 9<br />

��<br />

dyn/cm. 33,34 On the basis of the infinite-molecular-weight surface energies of the copolymer<br />

constituents (included in Table A3.1), addition of a SEBS block copolymer to an iPP/PLA<br />

blend lowers the interfacial tension at both the iPP-EB interface (3-5 dyn/cm) and the PS-<br />

PLA interface (~1 dyn/cm). Although Antonoff’s rule is overly simplistic for many systems,<br />

it has been shown 35 to accurately describe the interfacial energy between PS and PLA. The<br />

interfacial energy provides a measure of thermodynamic incompatibility between two<br />

dissimilar species and contributes to the enthalpic portion of the system free energy. As such,<br />

it relates 36,37,38 directly to the Flory-Huggins � interaction parameter and, by extension, to the<br />

corresponding difference in solubility parameters between the two species. The solubility<br />

parameters of all polymer species of interest here are also listed for comparison in Table<br />

A3.1. Solubility parameters are often used to estimate the compatibility of two polymers<br />

insofar as the species are relatively non-polar and mixing is endothermic. While there is no<br />

general guideline for predicting polymer-polymer miscibility from solubility parameters<br />

alone, 30 polymer pairs with nearly identical solubility parameters are more likely to be<br />

271


mutually miscible than those with even modestly different solubility parameters, regardless<br />

of their chemical constitution. 39 On the basis of their solubility parameters, the EB midblock<br />

of the copolymer should be compatible with iPP, whereas the S endblocks of the copolymer<br />

are not expected to show much preference for either iPP or PLA. We cannot comment much<br />

on the compatibility of the midblock extender in the copolymer, but it stands to reason that,<br />

since it is mixed with the EB midblock, it, too, will be compatible with iPP.<br />

To discern if the SEBS copolymer compatibilizes iPP/PLA blends, we examine the<br />

morphologies of blends composed of 50 wt% iPP. Figure A3.3a displays a cross-sectional<br />

SEM image of a melt-mixed iPP/PLA blend after removal of the dispersed PLA phase upon<br />

selective solvent exposure in DCM at ambient temperature. The diameter of the PLA<br />

domains (appearing as pores) is measured to be 5.0 ± 2.9 μm. Replacing 5 wt% of the PLA<br />

with SEBS is found to reduce the PLA domain diameter to 1.4 ± 1.3 μm (cf. Figure A3.3b),<br />

which confirms that the copolymer effectively compatibilizes the iPP/PLA blends, as<br />

anticipated from the thermodynamic considerations discussed above. It immediately follows<br />

that, if the copolymer molecules can locate along the interface between iPP and PLA, they<br />

should reduce slip (if it exists) during extrusion and improve the adhesion between iPP and<br />

PLA. It is of interest to note here that dissolution of the precompounded SEBS in DCM at a<br />

concentration of 5 wt% results in the formation of a cloudy solution that appears to remain<br />

stable to the unaided eye for several months at ambient temperature. We believe that the<br />

observed solution opacity arises from the self-organization of copolymer molecules into<br />

swollen micelles (spherical microemulsions) or vesicles 40 that are sufficiently large to scatter<br />

272


light and capable of encapsulating the midblock extender, which is most likely incompatible<br />

with DCM. A detailed account of this solution nanostructure is, however, beyond the scope<br />

of the present work.<br />

Multilayer Melt Rheology<br />

Various polymer pairs such as PS/poly(methyl methacrylate), PS/HDPE, PS/iPP,<br />

PE/fluoropolymer, and iPP/nylon-6 exhibit incompatibility-induced slip during rheological<br />

testing of layered specimens. 8,9,11 Layered specimens are more representative of the present<br />

bicomponent fibers than are blends, because the two polymer species are melt-contacted<br />

along an artificially-introduced interface rather than along multiple interfaces that develop in-<br />

situ due to thermodynamic instability. Generally speaking, "slip," or negative viscosity<br />

deviation, tends to increase with increasing polymer-polymer incompatibility. To determine<br />

whether iPP and PLA undergo slip during extrusion, rheological testing has been conducted<br />

on alternating multilayers of iPP and PLA. As illustrated in Figure 4, these sandwich<br />

structures are first positioned on the bottom flat plate of the rheometer, and then the top plate<br />

of the rheometer is lowered until it contacts the top polymer layer. We have measured the<br />

shear viscosity of molten polymer multilayers as a function of interfacial area, which is<br />

controllably manipulated by increasing the number of alternating layers, as well as the size of<br />

the plates. The zero-shear viscosities of these multilayers are presented as a function of<br />

interfacial area in Figure A3.4. To put these results into perspective, a melt-spun<br />

bicomponent fiber with an average inner diameter of 20 μm and measuring 1 m long would<br />

possess an interfacial area of approximately 0.6 cm 2 . Despite the greater incompatibility<br />

273


etween iPP and PLA as compared to other polymer pairs, no significant decrease in<br />

viscosity is detected with increasing interfacial area up to an interfacial area of ~90 cm 2 ,<br />

thereby indicating that slip between iPP and PLA is not detected by steady shear rheology at<br />

stresses below ~1 kPa.<br />

One reason for the absence of slip in this work may be due to the low shear rates<br />

accessible with a parallel-plate rheometer relative to capillary rheometers or fiber extrusion.<br />

Slip first becomes apparent, for instance, at interfacial areas between 34 and 152 cm 2 at shear<br />

stresses above 1 kPa for multilayers composed of iPP and PS. 8 Inertial forces cause the<br />

polymer to flow from the plates at stresses above 1 kPa, thereby resulting in poor data<br />

quality. Smaller parallel plates exacerbate this problem. Alternatively, we consider the<br />

manner by which previous studies concluded the existence of slip. While viscosities are<br />

reported, 8,9 the deviation from predicted "no-slip" conditions is often small. Due to the large<br />

variation in viscosity encountered in each sample, we have chosen to repeat each<br />

measurement 7 times in this work. Jiang et al. 9 have reported the viscosity of a PS/HDPE<br />

bilayer with an interfacial area of about 5 cm 2 to be 11% lower than the reciprocal rule of<br />

mixtures (R-ROM). On the basis of this comparison, they conclude this deviation is<br />

representative of slip. To discern the validity of this criterion, we apply several rules of<br />

mixtures to the data in Figure 4. The first is the R-ROM, which has been used to predict the<br />

viscosity of multilayered specimens composed of 2 polymers (indicated by subscripts 1 and<br />

2). It is given by<br />

��<br />

1<br />

� � �1 �<br />

�1 �2 �<br />

�2 �i �i 274<br />

(3)


where the subscripted i corresponds to the interfacial layer, which has few polymer<br />

entanglements and potentially lower viscosity. 41 In Eq. 3, � is the measured viscosity of the<br />

multilayer, , �1 and �2 are the volume fractions of polymers 1 and 2, respectively, and �1 and<br />

�2 are the viscosities of polymers 1 and 2, respectively, measured individually. Since no<br />

difference is evident in Figure 4 up to an interfacial area of ~90 cm 2 , this interfacial layer<br />

contribution is neglected.�<br />

Alternatively, other rules of mixtures should be considered. For example, the standard rule<br />

of mixtures (S-ROM) predicts viscosity according to<br />

��<br />

� � � 1 � 1 � � 2 � 2<br />

Moreover, a logarithmic rule of mixtures (L-ROM) has also been proposed for polymer<br />

blends, viz,<br />

14 14 14 42<br />

log� � � 1 log� 1 � � 2 log� 2 (5)<br />

Other rules include additional terms that attempt to explain deviations from the three<br />

provided in Eqs. 3-5<br />

��<br />

and are not considered here. The multilayer viscosity averaged over all<br />

the interfacial areas in Figure 4 is 754 Pa-s, whereas those predicted by Eqs. 3-5 are 714 (R-<br />

ROM), 746 (S-ROM) and 729 Pa-s (L-ROM). Thus, the S-ROM most closely predicts the<br />

viscosity of our multilayered samples, while the R-ROM is found to give the poorest<br />

prediction of the three. The standard errors of the PLA and iPP viscosities used in the ROM<br />

predictions are 9 and 3%, respectively, which suggests that measured multilayer viscosities<br />

deviating modestly (by 11% according to Jiang et al. 15 ) from predicted values are not<br />

statistically significant in systems composed of iPP and PLA. The relatively large variation in<br />

275<br />

(4)


measured multilayer and pure-component viscosities encountered here is indicative of the<br />

difficulty in confirming the existence of interfacial slip between iPP and PLA on a parallel-<br />

plate rheometer.<br />

Fiber Mechanical Properties<br />

The interface, which is largely responsible for the mechanical properties of bicomponent<br />

fibers (as well as fibers derived from polymer blends), is known 4 to fail during the<br />

mechanical drawing of bicomponent fibers spun from iPP and PLA. To reduce the inherent<br />

incompatibility between iPP and PLA, the SEBS copolymer is compounded with PLA before<br />

bicomponent fibers are co-spun. When PLA with and without copolymer is spun as the core<br />

(Figure A3.5a), the tenacity of the bicomponent fibers increases with increasing aspirator<br />

pressure up to a level near 2 cN/dtex. Incorporation of the copolymer has no systematic<br />

effect, positive or negative, on these results. Recall from Figure A3.1 that the PLA and iPP<br />

are both semi-crystalline in this fiber configuration. If the fibers are spun with an iPP core<br />

and PLA sheath, a similar dependence on aspirator pressure is seen (Figure A3.5b), but the<br />

maximum level attained is slightly lower for fibers with PLA only and lower still (but within<br />

experimental uncertainty) for fibers with a mixture of PLA+SEBS. In this configuration, the<br />

PLA is semi-crystalline, but the iPP is, for the most part, mesomorphic. Recall that, while the<br />

EB midblock of the SEBS copolymer is compatible with iPP, the S endblocks are not<br />

expected, on the basis of solubility parameters alone, to be strongly iPP- or PLA-compatible.<br />

Blending incompatible polymers into bicomponent fibers is expected to decrease fiber<br />

strength due to enlarged interfacial area and reduced interfacial strength. Yet, addition of the<br />

276


SEBS copolymer to PLA as the core (Figure A3.5a) or sheath (Figure A3.5b) does not<br />

compromise the mechanical properties of the fibers, indicating that the copolymer (with<br />

midblock extender) and PLA are not strongly incompatible.<br />

At the lowest aspirator pressure (fiber spinning velocity) in Figure A3.6, addition of<br />

copolymer into the PLA sheath is observed to increase substantially the strain at which<br />

rupture of the sheath occurs. As the aspirator pressure is increased, however, the strain at<br />

which PLA ruptures does not appear to be dependent on copolymer addition. Note that not all<br />

fibers undergo failure of the PLA sheath at lower strains than the iPP core. Rather, in these<br />

cases, a single catastrophic failure signals nearly simultaneous rupture of both polymers.<br />

Such events are accompanied by an abrupt increase in the true strain at break of the fibers.<br />

By adding the SEBS copolymer to PLA, the frequency at which the PLA sheath ruptures<br />

prior to failure of the iPP core is reduced at most aspirator pressures (cf. Figure A3.6), which<br />

indicates that the SEBS-modified PLA is capable of undergoing greater extension prior to<br />

failure. In the event that the PLA sheath does not fail before the iPP core (i.e., the sheath and<br />

core co-rupture) in any of the tested specimen, the frequency in Figure 6 is shown as 0%.<br />

While the Tg (Figure A3.7) and melt viscoelasticity (Figure A3.8) of PLA are not greatly<br />

affected by the incorporation of SEBS, the elastomeric nature of the copolymer, coupled with<br />

its intrinsic ability to self-organize into nanostructural elements, may serve to improve the<br />

elasticity of the PLA sheath, which, in turn, promotes an increase in the strain at which the<br />

PLA sheath ruptures. In Figure A3.7, addition of the SEBS copolymer alters the Tg of PLA<br />

by just over 1°C, which is considered to be within instrumental uncertainty. The melt<br />

277


frequency (�) spectra displayed in Figure A3.8 show little variation upon addition of the<br />

SEBS copolymer to PLA. In both cases, the dynamic loss modulus (G") scales as � 1.0 in the<br />

terminal region, which agrees with the behavior of entangled homopolymers (� 1 ). The<br />

analogous scaling behavior of the dynamic storage modulus (G'), which is � 2 for entangled<br />

homopolymers, changes upon copolymer addition from � 1.9 to � 1.6 . While this subtle<br />

reduction in slope may signify a copolymer-induced change in the molten structure, it is<br />

reasonable to expect that such a change might be more apparent in the solid state (where the<br />

styrenic endblocks are glassy) than in the melt. The copolymer-induced reduction in the<br />

frequency of PLA rupture at low strains may also indicate improved adhesion along the<br />

iPP/PLA interface, which would promote greater stress transfer to the iPP core and which<br />

would permit the PLA to support a higher stress before rupturing.<br />

Previously, some of us have introduced 4 a facile optical method for estimating the<br />

molecular orientation of bicomponent core/sheath fibers from true stress-true strain curves<br />

when the refractive indices of the core and sheath polymers significantly differ. Figure A3.9<br />

shows the progression of matching a fiber with an iPP core and PLA sheath to liquids of<br />

known refractive index so that the birefringence of the sheath, which is related to molecular<br />

orientation and fiber strength, can be determined. These images confirm that the conformal<br />

PLA sheath completely wets the iPP core. The liquid surrounding the fiber must possess a<br />

refractive index (RI) that is similar to those of the two species in the fiber to yield clear<br />

fringes (dark bands) that can be followed through each interface. Because the refractive<br />

indices of the iPP core and PLA sheath are considerably different (1.504 and 1.542,<br />

278


espectively, at 20°C for unoriented samples), no single liquid can provide clear fringes for<br />

both the core and sheath simultaneously, which explains why the fringes blur at the core-<br />

sheath and/or sheath-liquid interface in each of the images in Figure A3.9. In addition, while<br />

the interface of unstrained fibers appears continuous for the fibers in this study, birefringence<br />

measurements would be difficult or altogether impossible for strained iPPcore/PLAsheath fibers<br />

exhibiting voids at the polymer-polymer interface. Therefore, using the method described<br />

elsewhere, 4 we estimate the molecular orientation of PLAcore/iPPsheath fibers with and without<br />

added SEBS copolymer using stress-strain curves.<br />

The strain shift of single-component and bicomponent fibers, which relates proportionally<br />

to molecular orientation, is presented as a function of spinning velocity in Figure A3.10.<br />

Here, spinning velocity, rather than aspirator pressure, is used to facilitate comparison of<br />

birefringence from fibers produced by other spinning methods, such as melt spinning or<br />

electrospinning, that do not require an aspirator pressure. Bicomponent fibers exhibit<br />

markedly increased strain shift relative to iPP and PLA alone, and addition of the SEBS<br />

copolymer to the PLA core does not inhibit molecular orientation in the iPP sheath. This<br />

observation is consistent with previous studies 4,6 reporting that the sheath component dictates<br />

flow mechanics for core/sheath fiber extrusion, as well as the ultimate properties of<br />

bicomponent fibers, since the sheath component experiences the greatest stress in both the<br />

melt and solid phases. Spinning with an iPP sheath and a PLA (or PLA+SEBS) core serves to<br />

focus the spinline stress over a smaller cross-sectional area, which results in higher molecular<br />

orientation in the iPP sheath. While the sheath component may control flow mechanics, the<br />

279


core polymer is crucial to achieve the synergistic effect of enhanced molecular orientation in<br />

the sheath polymer beyond that which can be realized by spinning either polymer<br />

individually. The master stress-strain curve for PLA required in this methodology has been<br />

developed for PLA alone, not with SEBS copolymer. Therefore, we do not report strain shifts<br />

for iPPcore/PLAsheath fibers with and without SEBS to avoid attributing effects introduced by<br />

SEBS to changes in PLA molecular orientation.<br />

Nonequilibrium Copolymer Morphologies<br />

The morphologies of triblock copolymers, such as the SEBS copolymer employed in this<br />

study, have been the subject of numerous studies. Most commercial triblock copolymers are<br />

designed as thermoplastic elastomers with dispersed glassy microdomains embedded in, and<br />

connected to, a continuous, rubbery matrix. 46 Addition of a midblock-selective oil, 47<br />

tackifying resin 48 or homopolymer 49 to a triblock copolymer can yield the same<br />

morphologies observed in diblock copolymer/homopolymer blends, 50,51 in which case<br />

comparable design rules can be sensibly presumed. It immediately follows that, on the basis<br />

of the copolymer composition, the proprietary midblock extender added to the copolymer<br />

used here promotes a spherical or cylindrical morphology if the extender is largely or<br />

marginally miscible, respectively, with the EB midblock. Incorporation of the copolymer into<br />

PLA adds another level of complication, as the copolymer must now partition between the<br />

midblock extender and PLA, as well as interact with iPP along the iPP-PLA interface. Under<br />

ideal equilibrium conditions of slow solvent evaporation or melt mixing, followed by<br />

extensive solvent or thermal annealing, the resultant copolymer morphology may be<br />

280


complex, depending on the individual strengths of 6 different binary interactions (assuming<br />

that the EB midblock and midblock extender can each be treated as a single species). If the<br />

rapid melt processing of the bicomponent fibers modified by the copolymer is now<br />

considered, highly nonequilibrium morphologies can be reasonably expected.<br />

We begin with an overview of the morphologies of the SEBS copolymer in different<br />

fibers. Figure A3.11 shows a series of TEM images acquired from fibers varying in<br />

configuration, but spun at the same aspirator pressure (15 psi). In Figure A3.11a, a relatively<br />

low-magnification image of a bicomponent fiber composed of an iPP core and PLA sheath<br />

demonstrates that (i) the conformal Au coating around the edge of the fiber prevents swelling<br />

of the fiber with epoxy resin, which reacts with the vapor of RuO4(aq); (ii) the iPP-PLA<br />

interface is clearly differentiated in cross-section; and (iii) the iPP core is lightly stained by<br />

the vapor of RuO4(aq). An image of a PLA fiber containing 5 wt% SEBS copolymer is<br />

provided in Figure 11b and reveals that the copolymer molecules, selectively stained with the<br />

vapor of RuO4(aq), are present in the form of discrete and aperiodic nanostructures that most<br />

closely resemble dispersed tubules. The existence of tubules suggests that the midblock<br />

extender is either not highly compatible with the EB midblock of the copolymer or present at<br />

sufficiently high concentration to preclude appreciable solubilization within the copolymer<br />

matrix. Moreover, assuming that the copolymer is uniformly dispersed in the PLA prior to<br />

melt spinning, the styrenic endblocks of the copolymer do not appear to be very compatible<br />

with PLA.<br />

281


In the TEM images displayed in Figures A3.11c and A3.11d, the copolymer morphologies<br />

formed in bicomponent iPPcore/(PLA+SEBS)sheath and (PLA+SEBS)core/iPPsheath fibers are<br />

evident. In both cases, the copolymer forms dispersed nanostructures that remain distributed<br />

throughout PLA rather than accumulating along the iPP-PLA interface. This tendency is<br />

consistent with the property measurements provided earlier and explains why the copolymer<br />

does not significantly alter the breaking strength of the bicomponent fibers investigated here.<br />

Although thermodynamic considerations indicate that the copolymer should migrate to the<br />

iPP-PLA interface, the timescale associated with fiber spinning is on par with or faster than<br />

that required for the diffusion of individual copolymer molecules from the PLA phase to the<br />

interface. To complicate matters further, the copolymer molecules are assembled into<br />

nanostructures that are less driven (due to a concentration gradient) and slower to diffuse to<br />

the interface where they are needed to compatibilize the core and sheath. Careful<br />

examination of the nanostructures observed in Figures A3.11c and A3.11d reveals that the<br />

copolymer nanostructures tend to orient along a common direction (which may be normal to<br />

the iPP-PLA interface) and, more importantly, that the copolymer molecules appear to self-<br />

organize into tubules, not cylinders, in PLA co-spun with iPP. [Stained cylinders appear the<br />

most electron dense (darkest) along their centerline, whereas tubules are darkest along their<br />

periphery, due to thickness considerations in projection.] While bicomponent block<br />

copolymers have been previously reported 52,53 to form nanotubes, such morphologies<br />

normally require an additional driving force, such as crystallization, to do so.<br />

282


While it is intriguing that in all the core/sheath fiber configurations examined the<br />

copolymer molecules most often form tubules within PLA, more exotic morphologies, such<br />

as concentric tubules and spheres in tubules (what we term "peas in a pod"), are also<br />

observed, as evidenced by the TEM images provided in Figures A3.12a and A3.12b,<br />

respectively. Examples of the "peas in a pod" morphology are also highlighted in Figures<br />

A3.11c and A3.11d. A qualitatively similar sphere-in-cylinder morphology has been<br />

observed 54 in a thin film of an ABC triblock copolymer swollen with a good, neutral solvent.<br />

To the best of our knowledge, however, this morphology has not been previously reported for<br />

a blend of an ABA triblock copolymer dispersed in a C homopolymer. These highly non-<br />

classical morphologies are schematically depicted in Table A3.2, and relevant dimensions<br />

identified in the illustrations and measured from TEM images are included. To put these<br />

dimensions into perspective, the unperturbed gyration diameter of the blocks comprising the<br />

SEBS copolymer are calculated from the freely-jointed chain model 55 and the known block<br />

lengths. Since the EB midblock most likely adopts a looped or bridged conformation (which<br />

is rigorously true only at equilibrium), its molecular weight is halved so that it can be treated<br />

as a tail, a chain tethered at only one end, in similar fashion as the S endblocks. Block<br />

gyration diameters, estimated with the assumptions that (i) the statistical segment lengths of<br />

S and EB are comparable (~0.7 nm) 56,57 and (ii) the EB midblock consists of equal fractions<br />

of E and B, are ~4 nm for the S block and ~15 nm for (half) the EB block. The number of<br />

unperturbed blocks (N) corresponding to the measured dimensions identified in the diagrams<br />

shown in Table A3.2 is included in the same table and reveals several important features.<br />

283


Analysis of the tubule walls generally yields N ≈ 2, which corresponds to an endblock<br />

bilayer. This finding is consistent with the EB midblocks extending into both the tubular core<br />

and the surrounding matrix. In the case of a single tubule (morphology A in Table A3.2), the<br />

measured internal diameter also results in N ≈ 2 (i.e., 2 EB blocks) for tubules formed in the<br />

(PLA+SEBS)sheath configuration, but varies considerably from about 1 to 3 in the<br />

(PLA+SEBS)core configuration. This variation may reflect differences in the level of spin-line<br />

stress experienced by the copolymer molecules, as well as stochastic swelling of the EB<br />

midblock by the midblock extender or PLA. Similar variation is evident in the EB-rich<br />

regions of the concentric tubule morphology (morphology B in Table A3.2). Within the<br />

PLA+SEBS core, a single internal tubule appears highly swollen with N ≈ 8, but the distance<br />

between the internal and external walls yields N ≈ 1, which implies that the EB midblocks<br />

extending from the internal and external tubule walls are interdigitated into a monolayer,<br />

rather than bilayered. Lastly, in the "peas in a pod" morphology (morphology C in Table<br />

A3.2), the diameter of the internal spheres (which appear circular but, in a few cases, as<br />

shells) is significantly larger than two S endblocks (N ≈ 4). These enlarged spheres are<br />

presumed to be the result of partial PLA incorporation, although the possibility of other<br />

kinetically trapped species cannot be outright disregarded. The distances between<br />

neighboring spheres (which align along the direction of the tubule) and between the spheres<br />

and tubule wall in both the core and sheath fiber configurations yields N ≈ 1, which strongly<br />

suggests that the spheres are most likely connected together, as well as to the tubule walls, by<br />

bridged EB midblocks, as schematically depicted in Figure A3.13. Such molecular<br />

284


connectivity within the nanostructures present might serve to reinforce the PLA phase and<br />

improve its elasticity, as deduced from the flow and mechanical properties discussed earlier.<br />

It is comforting that these unique morphologies, although nonequilibrium in nature, tend to<br />

obey classical chain packing behavior.<br />

In addition to establishing the presence of uncommon copolymer morphologies in<br />

bicomponent fibers, the images presented in Figures A3.11b, A3.11c, A3.12a and A3.12b<br />

confirm that the S endblocks are not in direct contact with PLA. Rather, the EB midblock<br />

forms a contiguous coronal layer around the S-rich features. Such isolation helps to explain<br />

why the copolymer does not preferentially migrate to and accumulate along the iPP-PLA<br />

interface where it can promote compatibilization. To discern the extent to which the SEBS<br />

copolymer is dispersed within PLA prior to melt spinning, we have examined the melt-<br />

compounded PLA+SEBS mixture. Large structures such as those portrayed in Figure A3.12c<br />

are evident, indicating that the two materials are not thoroughly mixed. These<br />

macrostructures, measuring on the order of hundreds of nanometers across, also appear<br />

tubular (the figure displays a central "ring" and what appears to be an end cap). Close<br />

examination of their walls reveals an organized copolymer nanostructure. Complementary<br />

inspection of the as-received copolymer with midblock extender (Figure A3.12d) confirms<br />

the existence of an irregular morphology that resembles the nanostructured walls of the<br />

macroscale tubules in Figure A3.12c. Thus, we conclude that the copolymer retains some of<br />

its as-received nanostructure after being melt-blended with PLA, indicating that the<br />

285


compounding temperature and mechanical mixing employed here were insufficient to<br />

molecularly disperse the copolymer within PLA.<br />

On a side note, it is strangely curious, though, that the stainable, but minor, styrenic<br />

component in the as-received compound with the midblock extender appears as the matrix in<br />

Figure A3.12d, which suggests that the morphology may be more complicated than styrenic<br />

spheres or cylinders as previous anticipated. Although this figure is representative of entire<br />

sections of the as-received copolymer, it is likewise possible that the as-received copolymer<br />

is heterogeneous at macroscopic length scales. Such macroscale heterogeneity may be largely<br />

responsible for the unique nanoscale tubular morphologies reported here in Figures A3.11<br />

and A3.12, as they could be the result of subjecting the parent morphologies in the<br />

heterogeneous PLA+SEBS mixture to very high shear and extensional flow through the<br />

spinneret so that they became distorted and rearranged into lower-energy nanostructural<br />

elements. This nonequilibrium formation mechanism seems more plausible than conventional<br />

self-organization of the copolymer from a disordered state and would help to explain the<br />

variety of nanostructures observed on the basis of local copolymer composition and<br />

diffusion.<br />

Conclusions<br />

The strategy of copolymer-induced blend compatibilization has been applied to melt-spun<br />

bicomponent fibers that bring together iPP and PLA in contact along a single interface<br />

separating the core from the sheath. Although the two polymers are incompatible, there is no<br />

evidence of measurable slip at the molten iPP-PLA interface according to steady-shear<br />

286


heology. While the addition of a SEBS copolymer to a melt-mixed iPP/PLA blend serves as<br />

a compatibilizer by reducing the size of PLA domains arising from macrophase separation,<br />

incorporation of the copolymer into PLA prior to co-spinning does not drastically improve<br />

the properties of bicomponent iPP/PLA fibers, which implies that the copolymer molecules<br />

are unable to concentrate along the iPP-PLA interface during spinning. Despite an absence of<br />

copolymer along the fiber interface, the strain at which PLA as the sheath ruptures increases<br />

sharply at low spinning pressure, and the number of fibers that undergo failure of the PLA<br />

sheath prior to failure of the iPP core is reduced by up to 30%, upon addition of copolymer.<br />

Thus, although the copolymer does not modify the iPP-PLA interface, it does affect the host<br />

PLA by improving its elasticity through the formation of unique copolymer nanostructures<br />

that include single tubular, concentric tubular and "peas in a pod" morphologies. Careful<br />

analysis of such unexpected morphologies confirms the existence of nanostructural<br />

dimensions capable of accommodating local connectivity through midblock bridging, which<br />

consequently allows these highly-elastic, EB-rich copolymer dispersions to rubber-toughen 58<br />

the PLA.<br />

Acknowledgments<br />

Financial support of this work has been provided by the Nonwovens Cooperative<br />

Research Center at North Carolina State University and the National Science Foundation<br />

through a Graduate Research Fellowship (K. E. R.). We thank Dr. Behnam Pourdeyhimi for<br />

insightful and fruitful discussions, as well as Christina Tang and Alina Higham for analytical<br />

assistance.<br />

287


Tables<br />

Table A3.1. Physical properties of the polymers employed in this study.<br />

Polymer<br />

Tm 29 (°C)<br />

288<br />

Zero-shear<br />

viscosity at<br />

185°C<br />

(Pa-s)<br />

Surface free<br />

energy 30 at 20°C<br />

(dyn/cm)<br />

Solubility<br />

parameter 31,32<br />

(cal/cm 3 ) 1/2<br />

Ipp 165 a 901 a 30.1 7.4<br />

PLA 161 a 591 a 39.3 12.1<br />

SEBS copolymer — 6980 a — —<br />

Atactic polystyrene — — 40.7 9.5<br />

Low-density polyethylene — — 35.3 8.0<br />

Polybutylene — — 33.6 7.8<br />

a property was measured in the present study.


Table A3.2. Dimensions (in nm) of the SEBS copolymer morphologies observed in melt-<br />

spun bicomponent fibers.<br />

A B C<br />

Morphology A Morphology B Morphology C<br />

Core/sheath fiber d t di do d ds s g<br />

configuration a (n) (n) (n) (n) (n) (n) (n) (n)<br />

PLA + SEBS 51 ± 31 8 ± 4 126 225 56 ± 16 16 ± 6 10 ± 4 17 ± 4<br />

in core (50) (50) (1) (1) (12) (32) (32) (32)<br />

PLA + SEBS 30 ± 10 8 ± 2 70 100 42 ± 15 14 ± 4 13 ± 4 16 ± 5<br />

in sheath (50) (50) (2) (2) (13) (56) (56) (56)<br />

a n denotes the number of features measured.<br />

289


Figure A3.1. Filament diameter and polymer morphology as functions of aspirator<br />

pressure and fiber configuration, denoted by core (c) and sheath (s). While all filaments are at<br />

least partially amorphous, only crystalline or otherwise ordered morphologies are identified<br />

by symbols if present: �-crystalline iPP (filled squares), mesomorphic iPP (filled triangles)<br />

and crystalline PLA (open squares). Filaments spun at the same aspirator pressure with<br />

different configuration are connected by solid lines (color-coded and labeled).<br />

290


Figure A3.2. Filament diameter as a function of spinning pressure for different fiber<br />

configurations: (a) iPP and iPP in the sheath, and (b) PLA and PLA in the sheath.<br />

Homopolymers, bicomponent filaments without copolymer and bicomponent filaments with<br />

copolymer are color-coded black, blue and red and are correspondingly labeled. Error bars<br />

indicate the standard error in the data from ~10 trials, and the solid lines serve to connect the<br />

data.<br />

291


Figure A3.3. Cross-fracture SEM images of melt-mixed blends: (a) 50/50 w/w iPP/PLA<br />

and (b) 50/47.5/2.5 iPP/PLA/SEBS. The PLA is selectively removed upon immersion in<br />

DCM for ~100 h at ambient temperature.<br />

292


Figure A3.4. Zero-shear viscosities of iPP/PLA multilayers in different test<br />

configurations: PLA on the bottom with either 25 ( ) or 40 ( ) mm plates, or iPP on the<br />

bottom with 25 mm plates ( ). Included are the individual viscosities of iPP ( ) and PLA (<br />

). The error bars denote the standard error in the data, and the dashed line corresponds to the<br />

S-ROM prediction. The uncertainty on the prediction is based on the standard deviation of<br />

the component iPP and PLA viscosities used to calculate the viscosity of the multilayers.<br />

Included are schematic illustrations of the iPP/PLA multilayers: (a) PLA (blue) on bottom,<br />

(b) iPP (red) on bottom and (c) 8 alternating layers.<br />

293


Figure A3.5. Dependence of tenacity of iPP/PLA filaments on aspirator pressure for two<br />

different fiber configurations identified by the location of the iPP: (a) sheath and (b) core.<br />

Open and filled symbols identify specimens with and without added SEBS copolymer,<br />

respectively. Error bars indicate the standard error in the data from ~10 trials.<br />

294


Figure A3.6. (left axis) True strain at which the sheath of iPPcore/PLAsheath bicomponent<br />

filaments with ( ) and without ( ) added SEBS copolymer fail as a function of aspirator<br />

pressure. (right axis, red) Frequency of PLA sheath failure prior to iPP core rupture with ( )<br />

and without ( ) added copolymer. The solid lines serve to connect the data of systems<br />

without copolymer.<br />

295


Figure A3.7. Series of DSC thermograms collected at a cooling rate of 10°C/min from<br />

iPPcore/PLAsheath bicomponent filaments with and without SEBS copolymer added to the PLA<br />

phase at two aspirator pressures (labeled). The location of the PLA Tg in each system is<br />

marked.<br />

296


Figure A3.8. Frequency (�) spectra of the dynamic moduli (G', circles; G", triangles)<br />

measured from PLA (filled symbols) and PLA blended with 5 wt% SEBS (open symbols) at<br />

185°C and a stress of 5 Pa (in the linear viscoelastic limit). Shown for comparison are the<br />

scaling relationships expected for G' and G" in the low-� (terminal) region for entangled<br />

homopolymers.<br />

297


Figure A3.9. Optical micrographs of PPcore/PLAsheath bicomponent filaments (see<br />

schematic diagram) digitally recorded under polarized light to measure the birefringence<br />

according to ref. [43]. The measured birefringence can be related to molecular orientation<br />

and, thus, mechanical properties, as described elsewhere. 44,45 The refractive index (RI)<br />

provided on each image corresponds to that of the liquid surrounding the fiber.<br />

298


Figure A3.10. Dependence of the strain shift of iPP ( ), PLA ( ) and PLAcore/iPPsheath (<br />

) filaments on spinning velocity. Included are results ( ) from bicomponent filaments into<br />

which 2.5 wt% SEBS copolymer is added to the PLA phase. The lines serve as guides for the<br />

eye.<br />

299


Figure A3.11. Cross-sectional TEM images of filaments in different configurations. In (a),<br />

an iPPcore/PLAsheath bicomponent filament coated in Au and embedded in epoxy (labeled) is<br />

shown. A PLA filament modified with 5 wt% SEBS copolymer is presented in (b). In (c) and<br />

(d), iPPcore/(PLA+SEBS)sheath and (PLA+SEBS)core/iPPsheath bicomponent filaments are<br />

displayed. The styrene-rich copolymer nanostructures appear dark due to selective staining.<br />

In (c) and (d), the arrowheads identify a unique nanostructure: tubules with internal spheres<br />

("peas in a pod").<br />

300


Figure A3.12. Series of TEM images showing unexpected morphologies of the SEBS<br />

copolymer after being compounded with PLA and melt-spun into bicomponent filaments. In<br />

(a), single tubules and concentric tubules are evident, whereas the "peas in a pod"<br />

morphology is clearly seen in (b). Large-scale structures that appear vesicular with ordered<br />

copolymer walls are visible in PLA melt-compounded with the SEBS copolymer prior to cospinning<br />

(c). The morphology of the as-received copolymer with midblock extender is<br />

provided for reference in (d).<br />

301


Figure A3.13. Schematic illustrations of the unexpected SEBS morphologies observed in<br />

melt-spun bicomponent filaments composed of iPP and PLA: (a) single tubules with<br />

bilayered S (red) walls, (b) tubules swollen with either the midblock extender or entrapped<br />

PLA (green), (c) concentric tubules that are connected by bridged EB midblocks (blue), (d)<br />

concentric tubules swollen with either the midblock extender or entrapped PLA, (e) equally<br />

sized and spaced S spheres in a single tubule ("peas in a pod"), and (f) equally sized and<br />

spaced hollow S spheres in a single tubule.<br />

302


References<br />

[1] Kabeel, M. A. Rev. Sci. Instr. 1991, 62, 2950.<br />

[2] Sun, C.; Zhang, D.; Liu, Y.; Xiao, J. J. Ind. Textiles 2004, 34, 17.<br />

[3] Fedorova, N.; Pourdeyhimi, B. J. Appl. Polym. Sci. 2007, 104, 3434.<br />

[4] Park, C. -W. AIChE J. 1990, 36, 10.<br />

[5] Holsti-Miettinen, R.; Seppälä, J.; Ikkala, O. T. Polym. Eng. Sci. 1992, 32, 868.<br />

[6] Arvidson, S. A.; Wong, K. C.; Gorga, R. E.; Khan, S. A. Polymer (under review).<br />

[7] Okamoto, M.; Mizuguchi, S.; Watanabe, K. U. S. Patent #3,705,226, December 5, 1972.<br />

[8] Dugan, J. "Critical factors in engineering segmented bicomponent fiber for specific end<br />

uses." 1999 (http://www.FITfibers.com/Publications.htm).<br />

[9] Fruedenburg and Co Kg. (http://www.evolon.com).<br />

[10] Kikutani, T.; Radhakrishnan, J.; Arikawa, S.; Takaku, A.; Okui, N.; Jin, X.; Niwa, F.; Kudo,<br />

Y. J. Appl. Polym. Sci. 1996, 62, 1913.<br />

[11] Cho, H. H.; Kim, K. H.; Kang, Y. A.; Ito, H.; Kikutani, T. J. Appl. Polym. Sci. 2000, 77,<br />

2254.<br />

[12] Houis, S.; Schmid, M.; Lubben, J. J. Appl. Polym. Sci. 2007, 106, 1757.<br />

[13] Im, J. N.; Kim, J. K.; Kim, H. K.; Lee, K. Y.; Park, W. H. J. Biomed. Res. B 2007, 83B, 499.<br />

[14] Zhao, R.; Macosko, C. W. J. Rheol. 2002, 46, 145.<br />

[15] Jiang, L.; Lam, Y. C.; Yue, C. Y.; Tam, K. C.; Li, L.; Hu, X. J. Appl. Polym. Sci. 2003, 89,<br />

1464.<br />

[16] Lam, Y. C.; Jiang, L.; Yue, C. Y.; Tam, K. C.; Li, L. J. Rheol. 2003, 47, 795.<br />

[17] Park, H. E.; Lee, P. C.; Macosko, C. W. J. Rheol. 2010, 54, 1207.<br />

303


[18] Robeson, L. M. Polymer Blends: A Comprehensive Review. Hanser Gardner: Cincinnati,<br />

2007.<br />

[19] Molau, G. E. J. Polym. Sci. A 1965, 3, 1267.<br />

[20] Wang, D.; Li, Y.; Xie, X. M.; Guo, B. H. Polymer 2011, 52, 191.<br />

[21] Del Castillo-Castro, T.; Castillo-Ortega, M. M.; Herrera-Franco, P. J.; Rodriguez-Felix, D. E.<br />

J. Appl. Polym. Sci. 2010, 119, 2895.<br />

[22] Wei, B.; Genzer, J.; Spontak, R. J Langmuir 2004, 20, 8659.<br />

[23] Gozen, A. O.; Zhou, J.; Roskov, K. E.; Shi, A. -C.; Genzer, J.; Spontak, R. J. Soft Matter<br />

2011, 7, 3268.<br />

[24] Pernot, H.; Baumert, M.; Court, F.; Leibler, L. Nat. Mater. 2002, 1, 54.<br />

[25] Yoo, T. W.; Yoon, H. G.; Choi, S. J.; Kim, M. S.; Kim, Y. H.; Kim, W. N. Macromol. Res.<br />

2010, 18, 583.<br />

[26] Godshall, D.; White, C.; Wilkes, G. L. J. Appl. Polym. Sci. 2001, 80, 130.<br />

[27] Lipscomb, G .G. Polym. Adv. Technol. 1994, 5, 14.<br />

[28] Arvidson, S. A.; Khan, S. A.; Gorga, R. E. Macromolecules 2010, 43, 2916.<br />

[29] Mark, J. E. (Ed.) Polymer Data Handbook, Oxford University Press: New York, 1999.<br />

[30] Sperling, L. H. Polymeric Multicomponent Materials: An Introduction. John Wiley & Sons:<br />

New York, 1998.<br />

[31] Small, P. A. J. Appl. Chem. 1953, 3, 71.<br />

[32] Hoy, K. L. J. Paint Technol. 1970, 42, 76.<br />

[33] Antonow, G. N. J. Chim. Phys. 1907, 5, 372.<br />

[34] Antonow, G. N. J. Chim. Phys. 1907, 5, 8.<br />

[35] Biresaw, G.; Carriere, C. J. J. Polym. Sci. B: Polym. Phys. 2002, 40, 2248.<br />

304


[36] Helfand, E.; Tagami, Y. J. Polym. Sci., Part B. 1971, 9, 741.<br />

[37] Helfand, E.; Tagami, Y. J. Chem. Phys. 1972, 7, 3592.<br />

[38] Ermoshkin, A. V.; Semenov, A. N. Macromolecules 1996, 29, 6294.<br />

[39] Gaylord, N. G. Chemtech 1976, 6, 392.<br />

[40] Discher, D.; Eisenberg, A. Science 2002, 297, 967.<br />

[41] Lyngaae-Jorgensen, J. K.; Thomsen, D. Int. Polym. Proc. 1988, 2, 15.<br />

[42] Utracki, L. A. Polymer Alloys and Blends: Thermodynamics and Rheology. Hanser: Munich,<br />

1989.<br />

[43] Rangasamy, L.; Shim, E.; Pourdeyhimi, B. J. Appl. Polym. Sci. 2011, 121, 410.<br />

[44] Treloar, L. Trans. of the Faraday Soc. 1941, 37, 84-97.<br />

[45] Ward, I. Proc. of the Phys. Soc. 1962, 80, 1176.<br />

[46] Holden, G.; Legge, N. R.; Quirk, R. P.; Schroeder, H. E. (Eds.) Thermoplastic Elastomers,<br />

2 nd Ed., Hanser: Munich, 1996.<br />

[47] Krishnan, A. S.; Roskov, K. E.; Spontak, R. J. in Advanced Nanomaterials (K.E. Geckeler<br />

and H. Nishide, Eds.). Wiley-VCH: Weinheim, 2010, pp. 791-834.<br />

[48] Krishnan, A. S.; Seifert, S.; Lee, B.; Khan, S. A.; Spontak, R. J. Soft Matter 2010, 6, 4331.<br />

[49] Kane, L.; Norman, D. A.; White, S. A.; Matsen, M. W.; Satkowski, M. M.; Smith, S. D.;<br />

Spontak, R. J. Macromol. Rapid Commun. 2001, 22, 281.<br />

[50] Winey, K. I.; Thomas, E. L.; Fetters, L. J. J. Chem. Phys. 1991, 95, 9367.<br />

[51] Matsen, M. W. Macromolecules 1995, 28, 5765.<br />

[52] Raez, J.; Manners, I.; Winnik, M. A. J. Am. Chem. Soc. 2002, 124, 10381.<br />

[53] Wang, X.; Wang, H.; Frankowski, D. J.; Lam, P. G.; Welch, P. M.; Winnik, M. A.;<br />

Hartmann, J.; Manners, I.; Spontak, R. J. Adv. Mater. 2007, 19, 2279.<br />

305


[54] Elbs, H.; Drummer, C.; Abetz, V.; Krausch, G. Macromolecules 2002, 35, 5570.<br />

[55] Dealy, J. M.; Larson, R. G. Structure and Rheology of Molten Polymers: From Structure to<br />

Flow Behavior and Back Again. Hanser: Munich, 2006.<br />

[56] Laurer, J. H.; Khan, S. A.; Spontak, R. J.; Satkowski, M. M.; Grothaus, J. T.; Smith, S. D.<br />

Lin, J. S. Langmuir 1999, 15, 7947.<br />

[57] O'Connor, K. M.; Pochan, J. M.; Thiyagarajan, P. Polymer 1991, 32, 195.<br />

[58] Zhang, W.; Wei, F. Y.; Chen, L.; Zhang, Y. in Proc. 2009 Int. Conf. Adv. Fibers Polym.<br />

Mater., National Science Foundation of China: Shanghai, 2009, pp. 207-209.<br />

306


APPENDIX IV<br />

Enhanced Biomimetic Performance of Ionic Polymer-Metal Composite<br />

Abstract<br />

Actuators Prepared with Nanostructured Block Ionomers*<br />

Ionic polymer-metal composites (IPMCs) represent an important class of stimuli-<br />

responsive polymers that are capable of bending upon application of an electric potential.<br />

Conventional IPMCs, prepared with Nafion ® and related polyelectrolytes, often suffer from<br />

processing challenges, relatively low actuation levels and back relaxation during actuation. In<br />

this study, we examine and compare the effects of fabrication and solvent on the actuation<br />

behavior of a block ionomer with a sulfonated midblock and glassy endblocks that are<br />

capable of self-organizing and thus stabilizing a molecular network in the presence of a polar<br />

solvent. Unlike Nafion ® , this material can be readily dissolved and cast from solution to yield<br />

films that vary in thickness and exhibit enormous solvent uptake. Cycling the initial chemical<br />

deposition of Pt on the surfaces of swollen films (the compositing process) increases<br />

theextent to which the electrodes penetrate the films, thereby improving contact along<br />

thepolymer/electrode interface. The maximum bending actuation measured from IPMCs<br />

*This chapter has been published in its entirety:<br />

PH Vargantwar, KE Roskov, TK Ghosh, RJ Spontak.. “Enhanced Biomimetic Performance of Ionic Polymer-<br />

Metal Composite Actuators Prepared with Nanostructured Block Ionomers.” Accepted to Macromolecular<br />

Rapid Communications 9/2011.<br />

307


prepared with different solvents is at least comparable, but is often superior, to that reported<br />

for conventional IPMCs, without evidence of back relaxation. An unexpected characteristic<br />

observed here is that the actuation direction can be solvent-regulated. Our results confirm<br />

that this block ionomer constitutes an attractive alternative for use in IPMCs and their<br />

associated applications.<br />

Introduction<br />

Electroactive polymers (EAPs) constitute a rapidly growing genre in the blossoming field<br />

of stimuli-responsive polymers [1] and afford a wide variety of important material-related<br />

advantages over their inorganic counterparts. [2] Some of these benefits include light weight,<br />

low cost, mechanical robustness, facile processability, and straightforward scalability.<br />

Macromolecules classified as EAPs are further categorized according to the mechanism by<br />

which they undergo actuation upon application of an electrical potential. In the case of<br />

electronic EAPs, the mode of actuation depends on whether an applied field either promotes<br />

attraction of compliant, oppositely-charged surface electrodes to compress a low-modulus<br />

polymer (the electrostatic mechanism) or induces a change in molecular polarization and,<br />

hence, lattice dimensions of a semi-crystalline polymer (the electrostrictive mechanism). [3]<br />

Examples of electrostatic and electrostrictive EAPs are dielectric elastomers [4] and<br />

ferroelectric polymers, [5] respectively. Actuation can likewise occur by an ionic mechanism<br />

wherein ionic species migrate from one electrode to another and, by doing so, induce a<br />

change in the size and/or shape of a solvated polymer. Although systems developed with<br />

carbon nanotubes [6] fall into this category, we only consider further here ionic polymer-metal<br />

308


composites (IPMCs), wherein ion transport is accompanied by solvent diffusion and<br />

sufficient nonuniform polymer swelling to trigger a bending motion. [7,8] Due to their inherent<br />

similarity to biological muscle, IPMCs are often referred to as artificial muscle, and their<br />

electrical actuation is considered biomimetic. They can be used in motion-control devices [9]<br />

(e.g., micropumps, grippers, camera apertures, and catheters) and are alternatively suitable<br />

for energy harvesting [10] and vibration sensing. [11] Since the transport of ionic species<br />

necessitates a solvent-swollen hydrophilic polymer, most conventional IPMCs consist of a<br />

polyelectrolyte such as Nafion ® , a sulfonated fluorocarbon-based copolymer. In this study,<br />

we report how, with appropriate preparation, a block ionomer can provide comparable, if not<br />

superior, electroactuation performance.<br />

Block ionomers derive from block copolymers, which have become increasingly<br />

ubiquitous in fundamental and technological endeavors due largely to their unique ability to<br />

self-assemble into soft nanostructures. [12] Within this broad classification, linear ABA<br />

triblock copolymers and higher-order multiblock copolymers with at least one midblock<br />

provide the added, and often desirable, benefit of forming a contiguous network consisting<br />

predominantly of bridged midblocks that physically connect the copolymer nanostructure. [13]<br />

In this study, we focus on a block ionomer derived from a network-forming copolymer in the<br />

presence of a midblock-selective, low-volatility polar solvent. Consider, for illustrative<br />

purposes, an ABA triblock copolymer preferentially solvated with a B-selective solvent<br />

under equilibrium conditions. In the limit of high solvent concentrations, copolymer<br />

molecules self-organize into spherical micelles, each with an A core and a B corona. [14]<br />

309


While the long-range order of these micelles may become compromised, the network<br />

connecting the micelles remains intact unless the copolymer concentration is reduced below<br />

the critical gel concentration. Within a midblock-solvated network, micelles composed of<br />

glassy or semi-crystalline endblocks serve as network-stabilizing cross-link sites, thereby<br />

making such elastomeric gels suitable for a wide range of applications, including dielectric<br />

elastomers, [15] pressure-sensitive adhesives [16] and microfluidic substrates. [17] Recently, we<br />

have provided [18] evidence to show that this soft nanostructure design can be used in<br />

conjunction with a midblock-sulfonated pentablock ionomer (PBI) to generate highly<br />

responsive and stable IPMCs. In this work, we more closely explore the performance of these<br />

IPMCs under different preparation conditions and in different solvents.<br />

The PBI employed here is a poly[p-t-butyl styrene-b-(ethylene-alt-propylene)-b-(styrene-<br />

co-styrenesulfonate)-b-(ethylene-alt-propylene)-b-p-t-butyl styrene] ionomer, the chemical<br />

structure of which is displayed in Fig. A4.1a. While previous studies [19] have shown that<br />

endblock-sulfonated styrenic triblock copolymers can yield IPMCs, it must be recognized<br />

that the endblocks in the current system are incompatible with polar solvents and therefore<br />

form the glassy microdomains required for network stabilization, as schematically depicted<br />

in Fig. A4.1b. In the presence of a polar solvent, the sulfonated midblock of each molecule<br />

becomes highly swollen in a solvent-rich matrix and, for the most part, adopts either a<br />

bridged or looped conformation. Under equilibrium conditions (achieved, for instance, by<br />

mixing the copolymer and solvent together, followed by increasing the temperature to form a<br />

solution and then decreasing the temperature to promote copolymer self-assembly),<br />

310


endblock-rich micelles surrounded by a rubbery shell uniformly disperse throughout the<br />

solvent-rich matrix. The present study, however, utilizes a nonequilibrium process strategy<br />

(detailed in the Experimental Section) to generate nanostructured mesogels, [20] which retain<br />

morphological features of the solvent-free copolymer. The existence of a bridged-midblock<br />

network connecting, and stabilized by, glassy endblock-rich microdomains can be confirmed<br />

by dynamic rh<br />

acquired at ambient temperature in the linear viscoelastic regime are shown in Fig. 1c for the<br />

PBI selectively swollen with two hydrophilic solvents used here: glycerol (GLY) and<br />

ethylene glycol (EG) at their maximum solvent uptake levels (~450 and ~500 wt% GLY and<br />

EG, respectively). Two signature features evident in this figure are consistent with a<br />

physically cross-linked network: [21] (i<br />

(confirming the load-bearing capability of the network), and (ii) G' is weakly (if at all)<br />

and nearly parallel to G".<br />

In general, IPMCs are fabricated from a polyelectrolyte membrane whose surfaces are<br />

coated with metals such as Pt, Au or Ag. [7,22,23] The IPMCs generated here begin with water-<br />

swollen PBI films that are subjected to Pt chemical deposition to metalize the inner surface,<br />

followed by surface electroding with Ag. Since the electrodes must establish a large contact<br />

area with the polyelectrolyte to achieve satisfactory capacitance, the electrode morphology<br />

constitutes a critical design consideration. Good interfacial adhesion between the electrode<br />

and polyelectrolyte is required for an efficient electric double layer, which, in turn, governs<br />

311


the storage capacity (and actuation strain) of IPMCs. [24] During conventional IPMC<br />

fabrication, a polyelectrolyte membrane is subjected to surface roughening and mechanical<br />

prestrain to improve the adhesion and distribution of the electrode metal along the<br />

polymer/metal interface. Since PBI swells extensively in deionized water at elevated<br />

temperatures and then collapses in the tetraamine platinum(II) chloride hydrate solution used<br />

for Pt electroding (due to salting-out), no such surface pretreatment is required in the present<br />

study. This feature, coupled with our ability to vary the number of Pt coatings applied,<br />

permits precise control over the distribution of Pt to desired depths within the membrane,<br />

which dictates the effective interfacial area and, in so doing, governs the actuation<br />

performance, as discussed later. On the basis of the findings reported by Lu et al., [25] we first<br />

apply one or more Pt inner-surface electrode layers, followed by an Ag surface electrode<br />

layer, [23] to improve actuation performance. The distribution of Pt in the IPMC after each Pt<br />

deposition cycle has been probed by scanning electron microscopy (SEM) and energy-<br />

dispersive spectroscopy (EDS), and results obtained after 3 consecutive cycles are presented<br />

in Figs. A4.2a-c. The cross-sectional SEM image displayed in Fig. A4.2a reveals that the PBI<br />

membrane is sharply delineated from the top and bottom electrode layers composed of Pt and<br />

Ag. The Pt layer is not readily observed after a single coating, but it becomes increasingly<br />

evident upon deposition cycling.<br />

Corresponding elemental x-ray maps acquired by EDS are shown in Figs. A4.2b-c for 3<br />

Pt deposition cycles and confirm the spatial distribution of Pt (Fig. A4.2b) and Ag (Fig.<br />

A4.2c). It is important to recognize that, although Ag is restricted to the outer surface, Pt<br />

312


migrates into the membrane upon deposition cycling. This aspect of the IPMC electrodes is<br />

more clearly seen in the EDS line scans provided for 1, 2 and 3 cycles in Fig. A4.2d. As the<br />

and the concentration of Pt within the membrane correspondingly increases. In contrast,<br />

electrode metals are found [26] to penetrate to only 10-20 μm in Nafion ® . The Ag surface<br />

electrodes measuring 30-40 μm thick reduce the surface resistance to 2 �, which ensures that<br />

the potential along the entire length of the actuator is relatively uniform during testing. Recall<br />

that the electrodes are sequentially applied to both surfaces of water-swollen IPMCs. If<br />

desired, the water can be replaced by other polar solvents, such as GLY and EG, that are<br />

miscible in water. Regardless of the nature of the solvent, the free acid form of the ionomer is<br />

exchanged with Li ions, which promote substantially increased solvation. Application of an<br />

electric potential across the thickness of the IPMC results in bending actuation, as illustrated<br />

by the schematic diagram presented in Fig. A4.1b and confirmed by the real-time sequence<br />

of superimposed images (discussed later) shown in Figs. A4.3a and A4.3b. While several<br />

mechanisms have been proposed [27] to explain such actuation, one of the most widely<br />

accepted models is based on the microelectromechanical (MEM) model proposed by Asaka<br />

and Oguro. [28] Recent neutron imaging and fluorescence microscopy studies performed by<br />

Park et al. [8] and Park et al., [7] respectively, provide direct experimental evidence of this<br />

mechanism in IPMCs composed of Nafion ® .<br />

The MEM mechanism requires a potential-driven spatial redistribution of counterions and<br />

solvent molecules within the nanostructured ionomer membrane of the IPMC. In the<br />

313


cantilever test configuration, the electric potential causes solvated cations to migrate towards<br />

the cathode, thereby increasing the electro-osmotic pressure of the ionic clusters along that<br />

side of the specimen. As solvent enters the clusters, the internal strain in the polymer matrix<br />

swells the interfacial boundary layer separating the cathode and membrane, and a<br />

solvent/counterion gradient forms across the membrane. Due to depletion of solvent<br />

molecules and counterions at the anode, the boundary layer separating the anode and<br />

membrane contracts. Concurrent swelling and contraction along opposite boundary layers<br />

and about the neutral plane results in the IPMC bending towards the anode. The rate at<br />

which, as well as the extent of which, bending occurs depends sensitively on several material<br />

characteristics, including the nanostructure and composition of the membrane, the quality of<br />

the solvent, and the nature of the neutralizing cation. The two most commonly employed<br />

polyelectrolytes used in IPMC actuators are Nafion ® (and chemically-related analogs) and<br />

endblock-sulfonated triblock copolymers. Unlike Nafion ® , [29] the PBI introduced here affords<br />

(i) facile processing in solvents without sacrificing useful mechanical or conductive<br />

properties, (ii) considerably higher solvent absorption levels, (iii) tunable morphologies in<br />

terms of nanostructural dimensions and anisotropy, and (iv) precise control over electrode<br />

morphology. Solvent processing can be used to tailor membrane dimensions, whereas solvent<br />

uptake determines the ionic conductivity of IPMCs and their actuation efficacy. Since most<br />

prior studies of IPMCs rely exclusively on the use of Nafion ® , control over membrane<br />

morphology provides a largely unexplored route by which to alter ion and solvent transport<br />

and, thus, the rate and extent of actuation. While endblock-sulfonated triblock copolymers<br />

314


have, in fact, been used [19] to a limited extent as IPMC membranes, introduction of a polar<br />

solvent plasticizes the glassy endblocks and compromises the mechanical properties of such<br />

IPMCs at the solvent concentrations needed for satisfactory conductivity. This trade-off is<br />

not encountered in the current PBI membrane design.<br />

The electromechanical performance of IPMC actuators is quantified by measuring<br />

cantilever deflection relative to zero load. In this work, measurements have been conducted<br />

at electric potentials ranging from 0.5 to 9 V, depending on the solvent employed. Although<br />

most of the results presented here correspond to organic (GLY and EG) solvents, water is<br />

used initially to permit comparison with other IPMCs, as shown in Fig. A4.3c. Bending<br />

deflections generated here with a dc potential of 1 V are marginally greater than those<br />

achieved with Nafion ® actuated with an ac potential of 1 V at 0.5 Hz. Note, however, that the<br />

PBI investigated here yields actuation levels that are more than 3x greater than an IPMC<br />

derived from an endblock-sulfonated triblock copolymer subjected to a dc potential of 1 V,<br />

and more than 6x greater than an IPMC fabricated from Nafion ® (exchanged with Na ions)<br />

under a dc potential of 2 V. [30] While the PBI membrane requires more time (~30 s) to<br />

achieve full actuation and the Nafion ® -based IPMC actuates to its maximum in ~5 s, the<br />

latter also shows considerably greater evidence of back relaxation, in which the IPMC<br />

subjected to continued electrical bias changes its actuation direction (i.e., towards the<br />

original, zero-load position) after reaching a maximum forward deflection. Back relaxation<br />

has been previously reported [26] to depend on the nature of the ionic moiety present on the<br />

polyelectrolyte backbone and molecular-level relaxation processes associated with the<br />

315


polymer chains comprising the membrane. Actuation tests using water as the conductive<br />

solvent are limited by the relatively low normal boiling point and electrolysis potential of<br />

water. To avoid these limitations and reduce the extensive likelihood of oxidizing the<br />

electrode metals, we have elected to use GLY and EG in the remainder of this study. Their<br />

normal boiling points are 290 and 197°C, respectively, whereas their corresponding<br />

electrolysis potentials are ~10 V for GLY and ~4 V for EG. One drawback to using organic<br />

liquids instead of water is that their viscosities are significantly higher (934 cP for GLY and<br />

16 cP for EG, relative to 1 cP for water, at 25°C), which is expected to result in markedly<br />

slower actuation rates.<br />

A sequence of digital images acquired during actuation of a GLY-containing IPMC<br />

fabricated from the PBI (with 3 Pt electrode coatings) and exposed to an electric potential of<br />

7 V is displayed in Fig. A4.3a. This sequence demonstrates that the active length (L) of the<br />

IPMC (not gripped at the clamps) undergoes significant bending deflection upon actuation,<br />

the extent of which is measured in terms of the dimensionless parameter �L, where � is the<br />

curvature (equal to the reciprocal of the radius of curvature) of the IPMC cantilever strip.<br />

Values of �L extracted from digital images such as those in Fig. 3a confirm that bending<br />

proceeds to a plateau level and that the magnitude of bending increases with (i) the number<br />

of Pt coatings applied and (ii) the applied potential up to 7 V. At higher potentials (9 V, not<br />

shown here), the actuation level decreases due presumably to observed loss of solvent<br />

through the nanoporous Ag electrode during actuation. Included for comparison in this figure<br />

are actuation kinetics of IPMCs produced with Nafion ® exchanged with different ionic<br />

316


species and actuated in GLY at 2 V. As in Fig. A4.3c, Nafion ® -based IPMCs actuate more<br />

quickly than those containing the PBI membrane, but the Nafion ® systems also exhibit<br />

considerable back relaxation while the potential remains applied. Back relaxation, which<br />

hinders control of actuator motion and peak force, is not evident in the PBI-containing<br />

IPMCs over the course of the actuation measurement, but relaxation back to zero deflection<br />

ensues upon removal of the potential. It is of further interest that the IPMCs fabricated from<br />

Nafion ® with K ions bend, for the most part, towards the cathode following a small initial<br />

deflection toward the anode. Similar behavior, however with no initial bending towards the<br />

anode, is observed with the PBI membrane (and Li ions) swollen with EG, as evidenced by<br />

the images shown in Fig. A4.3b and the extracted data presented in Fig. A4.3e.<br />

The sequence of images provided in Fig. A4.3b verifies that the IPMC undergoes<br />

substantial bending toward the cathode upon electrical actuation. As far as we are aware, this<br />

is the first time that the direction of bending actuation is solvent-regulated, holding all other<br />

parameters constant. Although the precise reason for this unexpected behavior is not known<br />

at this time, we hypothesize that it is a consequence of the chemical interaction between EG<br />

molecules and the functional groups on the PBI chains or the mobile Li ions, and is not due<br />

to viscosity differences. As with GLY in Fig. A4.3d, the extent of actuation increases to a<br />

plateau level as the electrical potential is increased to 1.0-1.5 V, beyond which a reduction in<br />

bending is observed most likely for the same reasons offered to explain the maximum voltage<br />

encountered with GLY. When the voltage is removed, no back relaxation occurs, further<br />

confirming that solvent quality plays a critical role in the fabrication of PBI-containing<br />

317


IPMCs with targeted actuation properties. To compare the transport properties of the data<br />

presented in Figs. A4.3d and A4.3e, we apply the MEM framework [28] to account for bending<br />

as a function of time (t), but recognize that other models are likewise available for this<br />

purpose. According to this theory (originally developed for water-swollen IPMCs), the<br />

curvature of an initially flat membrane can be conveniently written as<br />

� � �<br />

1� exp ��<br />

Dm 2 Dmt /W 2 � ( ��<br />

(1)<br />

where � for a given specimen under actuation is a constant that depends on several<br />

parameters, �� including the thickness (W), equilibrium solvation and modulus of the<br />

membrane, as well as the diffusivity (Dm), density and transfer coefficient of the solvent, and<br />

the applied current density.<br />

Regression of Eq. 1 to the data yields the solid lines included in Figs. A4.3d and A4.3e and<br />

the values of Dm provided as a function of applied voltage in Fig. A4.3f. For comparison, the<br />

value of Dm for water in a Nafion ® -based IPMC at 25°C is ~4.5 x 10 -6 cm 2 /s, [28] whereas<br />

those for GLY and EG in the PBI tend to be lower, ranging from about 3 x 10 -6 cm 2 /s for EG<br />

to 1 x 10 -7 cm 2 /s for GLY.<br />

Maximum values of �L measured from this study are compared with those from previous<br />

efforts in Fig. A4.4 and immediately reveal that IPMCs containing the PBI membrane<br />

swollen with GLY are capable of substantially larger actuation levels at higher voltages. This<br />

figure also provides insight into the dependence of actuation on the number of Pt deposition<br />

cycles. We hasten to point out that if the number of Pt cycles is increased to 4, the actuation<br />

properties deteriorate as a percolated network of Pt develops and permits uninterrupted<br />

318


current flow. In addition, the results displayed in this figure confirm that the direction of<br />

actuation can be controlled by judicious selection of the solvent employed. Thus,<br />

incorporation of the PBI membrane into an IPMC permits tunable bending actuation and<br />

direction at high solvent concentrations without sacrificing the mechanical integrity of the<br />

IPMC. The novel multicomponent design afforded by the PBI membrane can provide<br />

unprecedented access to next-generation IPMCs with tailorable morphologies and, hence,<br />

ion/solvent transport properties and (electro)mechanical properties for motion-control<br />

devices, designer sensors and bio-inspired applications such as artificial cilia [31] and<br />

wormlike robots. [32]<br />

Experimental<br />

The PBI used here was generously supplied by Kraton Polymers (Houston, TX). The<br />

sizes of the blocks (cf. Fig. A4.1a) in the non-sulfonated parent copolymer were<br />

approximately 15 (tbS), 10 (EP) and 28 (S) kDa. As reported by the manufacturer, the<br />

midblock was about 57 mol% sulfonated, and the ion exchange capacity (IEC) of the PBI<br />

was 2.0 meq/g polymer. Reagent-grade tetrahydrofuran (THF), GLY, EG, and lithium<br />

chloride were purchased from Fisher Scientific and used as-received. Tetraamine<br />

platinum(II) chloride hydrate was obtained from Aldrich. The PBI was dissolved in THF to<br />

form 2 wt% solutions, which were subsequently cast into Teflon molds to form films that<br />

were annealed at 60°C for 24 h under vacuum. These films, measuring ~400 μm thick, were<br />

swollen with GLY or EG at 60°C overnight, and the resulting materials were analyzed on a<br />

Rheometrics Mechanical Spectrometer (RMS 800) equipped with 8 mm parallel plates<br />

319


separated by a 1-2 mm gap. The linear viscoelastic (LVE) limits at ambient temperature were<br />

determined by performing dynamic strain sweeps from 0.5 to 10% strain amplitude at a<br />

frequency of 1 rad/s. Frequency spectra were acquired at a strain amplitude of 1% in the LVE<br />

region. Fabrication of IPMCs required the formation of electrodes on the top and bottom<br />

surfaces of solvent-cast films. This involved two steps: initial compositing and surface<br />

electroding. In the first step, the top and bottom Pt electrodes were formed on a water-<br />

swollen PBI film by electroless deposition of Pt, as described elsewhere. [25] This procedure<br />

was repeated up to 4 times. Surface electroding resulted in the formation of Ag electrodes on<br />

top of the Pt electrodes by the silver mirror reaction. [23] As the reaction proceeded, discrete<br />

Ag particles deposited on the film surfaces and formed a contiguous layer. The dried<br />

membrane/electrode assembly thus obtained was examined by scanning electron microscopy<br />

(SEM) conducted in a Hitachi S3200 variable-pressure instrument operated at 20 kV in the<br />

presence of He at a pressure of 100 Pa. Samples were frozen in liquid nitrogen and cross-<br />

fractured to differentiate the layered morphology without the use of a conductive coating. An<br />

Oxford Isis EDS system was used to measure the spatial distribution of the electrode metals.<br />

For electromechanical characterization, fabricated IPMC films with electrodes were swollen<br />

in water, followed by immersion in a 0.5 M lithium chloride solution for 24 h at 60°C to<br />

promote Li ion exchange. Each film was subsequently submersed in a predetermined organic<br />

solvent at 60°C for 12 h to produce a mesogel (which retains morphological characteristics of<br />

the PBI prior to introduction of solvent) and then cut into pieces measuring 3 mm x 25 mm.<br />

One end of the cut strip was clamped, and a free (active) length (L) measuring 1.9 to 2.0 cm<br />

320


long was continuously monitored as a predetermined electric potential was applied through<br />

the clamps. Recorded digital footage was analyzed by the Matrox Inspector software to<br />

determine the level of bending actuation from the radius of curvature along L.<br />

Acknowledgments<br />

This study was funded by the National Science Foundation, and K. E. R. gratefully<br />

acknowledges support from a National Science Foundation Graduate Fellowship. We thank<br />

Professor S. A. Khan for use of his laboratory facilities.<br />

321


Figures<br />

Figure A4.1. (a) Chemical structure of the PBI block ionomer possessing p-t-butyl styrene<br />

[tbS] endblocks separated from the styrene-co-styrenesulfonate [S(sS)] midblock by flexible<br />

ethylene-alt-propylene [EP] linkages. (b) Schematic illustration of the idealized microphaseseparated<br />

morphology of a midblock-swollen PBI film — the copolymer shading matches<br />

that in (a) — and the resulting actuation mechanism of IPMCs prepared therefrom. (c)<br />

Frequency spectra of the PBI selectively swollen with GLY (open symbols) and EG (filled<br />

symbols). The dynamic storage (G', circles) and loss (G", triangles) moduli are labeled.<br />

322


Figure A4.2. (a) SEM cross-sectional image and corresponding elemental x-ray maps of (b)<br />

Pt and (c) Ag after 3 Pt deposition cycles. Line scans collected from cross-sectional x-ray<br />

maps for Pt (light line) and Ag (dark line) are included in (d) for IPMCs subjected to 1, 2 and<br />

3 Pt cycles.<br />

323


Figure A4.3. Digital image sequences of PBI-based IPMCs acquired during electroactuation<br />

with (a) GLY and (b) EG at 7 and 1 V, respectively. (c) Actuation-induced bending<br />

deflection of Li-exchanged, hydrated PBI ( ) at 1 V. For comparison, deflection results<br />

reported for hydrated IPMCs composed of an endblock-sulfonated triblock copolymer ( ) at<br />

1 V [19] and Li-exchanged Nafion® (dashed line) at 1 V and 0.5 Hz[19] and Na-exchanged<br />

Nafion® ( ) at 2 V[30] are included. (d) Bending actuation as a function of time for IPMCs<br />

consisting of GLY-solvated PBI (labeled with the cation and applied voltage) and Nafion®<br />

(labeled with the cation in parentheses) at 2 V.[26] (e) Bending actuation as a function of<br />

time for IPMCs consisting of EG-solvated PBI (labeled with the cation and applied voltage),<br />

as well as Nafion® ( , ) and Flemion® ( ) (each labeled with the cation and applied voltage<br />

in parentheses).[26] In both (d) and (e), the effect of Pt coating on bending actuation is<br />

shown for PBI-based IPMCs ( , labeled with the number of Pt cycles) at 7 and 1 V,<br />

respectively. The dashed lines in (d) and (e) serve to connect literature data, whereas the<br />

solid lines correspond to regressions of Eq. 1 to data collected here. The bending direction of<br />

the IPMC is identified by the background shading: anode (white) or cathode (gray). (f)<br />

Values of Dm extracted from the regression analyses in (d) and (e) and presented as a<br />

function of electric potential for PBI-based IPMCs solvated with GLY (circles) and EG<br />

(triangles) and subjected to 2 (filled) and 3 (open) Pt deposition cycles. The solid and dashed<br />

lines in (f) serve to connect the data.<br />

324


Figure A4.4. Maximum bending actuation achieved for IPMCs fabricated from the PBI<br />

investigated in this work. Shown here are the IPMCs with GLY (filled symbols) and EG<br />

(open symbols), each labeled with the number of Pt deposition cycles. Error bars correspond<br />

to the standard error. The cross-hatched region identifies the range of actuation levels<br />

achieved for conventional IPMCs under similar test conditions. Background shading is the<br />

same as in Fig. 3, and the solid lines serve to connect the data.<br />

325


References<br />

[1] a) A. K. Bajpai, J. Bajpai, R. Saini, R. Gupta, Polym. Rev. 2011, 51, 53; b) D. Roy, J. N.<br />

Cambre, B. S. Sumerlin, Prog. Polym. Sci. 2010, 35, 278.<br />

[2] a) E. Hornbogen, Adv. Eng. Mater. 2006, 8, 101; b) S. A. Wilson, R. P. J. Jourdain, Q. Zhang,<br />

R. A. Dorey, C. R. Bowen, M. Willander, Q. U. Wahab, M. Willander, M. A. H. Safaa, O.<br />

Nur, E. Quandt, C. Johansson, E. Pagounis, M. Kohl, J. Matovic, B. Samel, W. van der<br />

Wijngaart, E. W. H. Jager, D. Carlsson, Z. Djinovic, M. Wegener, C. Moldovan, R. Iosub, E.<br />

Abad, M. Wendlandt, C. Rusu, K. Persson, Mater. Sci. Eng. R. 2007, 56, 1.<br />

[3] a) T. Mirfakhrai, J. D. W. Madden, R. H. Baughman, Mater. Today 2007, 10, 30; b) R.<br />

Shankar, T. K. Ghosh, R. J. Spontak, Soft Matter 2007, 3, 1116; c) P. Brochu, Q. B. Pei,<br />

Macromol. Rapid Comm. 2010, 31, 10.<br />

[4] a) R. Pelrine, R. Kornbluh, Q. B. Pei, J. Joseph, Science 2000, 287, 836; b) F. Carpi, S.<br />

Bauer, D. De Rossi, Science 2010, 330, 1759.<br />

[5] a) Q. M. Zhang, H. F. Li, M. Poh, F. Xia, Z. Y. Cheng, H. S. Xu, C. Huang, Nature 2002,<br />

419, 284; b) J. J. Li, S. I. Seok, B. J. Chu, F. Dogan, Q. M. Zhang, Q. Wang, Adv. Mater.<br />

2009, 21, 217.<br />

[6] V. H. Ebron, Z. W. Yang, D. J. Seyer, M. E. Kozlov, J. Y. Oh, H. Xie, J. Razal, L. J. Hall, J.<br />

P. Ferraris, A. G. MacDiarmid, R. H. Baughman, Science 2006, 311, 1580.<br />

[7] I. S. Park, S. M. Kim, D. Pugal, L. M. Huang, S. W. Tam-Chang, K. J. Kim, Appl. Phys. Lett.<br />

2010, 96, 043301.<br />

[8] J. K. Park, P. J. Jones, C. Sahagun, K. A. Page, D. S. Hussey, D. L. Jacobson, S. E. Morgan,<br />

R. B. Moore, Soft Matter 2010, 6, 1444.<br />

326


[9] a) M. Shahinpoor, K. J. Kim, Smart Mater. Struct. 2005, 14, 197; b) A. J. Duncan, D. J. Leo,<br />

T. E. Long, Macromolecules 2008, 41, 7765.<br />

[10] J. Brufau-Penella, M. Puig-Vidal, P. Giannone, S. Graziani, S. Strazzeri, Smart Mater. Struct.<br />

2008, 17, 015009.<br />

[11] K. Krishen, Acta Astronaut. 2009, 64, 1160.<br />

[12] a) I. W. Hamley, The Physics of Block Copolymers, Oxford University Press, New York<br />

1998; b) M. Q. Li, C. K. Ober, Mater. Today 2006, 9, 30; c) M. C. Orilall, U. Wiesner, Chem.<br />

Soc. Rev. 2011, 40, 520.<br />

[13] a) M. W. Hamersky, S. D. Smith, A. O. Gozen, R. J. Spontak, Phys. Rev. Lett. 2005, 95,<br />

168306.; b) A. S. Krishnan, K. E. Roskov, R. J. Spontak, in Advanced Nanomaterials, (Eds: K. E.<br />

Geckeler, H. Nishide), Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany 2010, Ch. 26.<br />

[14] J. H. Laurer, J. F. Mulling, S. A. Khan, R. J. Spontak, R. Bukovnik, J. Polym. Sci. Part B<br />

1998, 36, 2379.<br />

[15] R. Shankar, T. K. Ghosh, R. J. Spontak, Adv. Mater. 2007, 19, 2218.<br />

[16] N. Jullian, L. Rubatat, P. Gerard, J. Peyrelasse, C. Derail, J. Rheol. 2011, 55, 379.<br />

[17] A. P. Sudarsan, J. Wang, V. M. Ugaz, Anal. Chem. 2005, 77, 5167.<br />

[18] P. H. Vargantwar, R. Shankar, A. S. Krishnan, T. K. Ghosh, R. J. Spontak, Soft Matter 2011,<br />

7, 1651.<br />

[19] X. L. Wang, I. K. Oh, J. Lu, J. H. Ju, S. Lee, Mater. Lett. 2007, 61, 5117.<br />

[20] M. R. King, S. A. White, S. D. Smith, R. J. Spontak, Langmuir 1999, 15, 7886.<br />

[21] Y. Y. He, T. P. Lodge, Chem. Commun. 2007, 2732.<br />

[22] N. Fujiwara, K. Asaka, Y. Nishimura, K. Oguro, E. Torikai, Chem. Mater. 2000, 12, 1750.<br />

[23] H. Tamagawa, K. Yagasaki, F. Nogata, J. Appl. Phys. 2002, 92, 7614.<br />

327


[24] M. Aureli, W. Y. Lin, M. Porfiri, J. Appl. Phys. 2009, 105, 104911.<br />

[25] J. Lu, S. G. Kim, S. Lee, I. K. Oh, Adv. Funct. Mater. 2008, 18, 1290.<br />

[26] S. Nemat-Nassera, S. Zamani, Y. Tor, J. Appl. Phys. 2006, 99, 104902.<br />

[27] a) Electroactive Polymer (EAP) Actuators as Artificial Muscles: Reality, Potential, and<br />

Challenges, 2nd ed. (Ed: Y. Bar-Cohen), SPIE, Bellingham, WA 2004; b) S. Nemat-Nasser,<br />

J. Y. Li, J. Appl. Phys. 2000, 87, 3321.<br />

[28] K. Asaka, K. Oguro, J. Electroanal. Chem. 2000, 480, 186.<br />

[29] M. A. Hickner, H. Ghassemi, Y. S. Kim, B. R. Einsla, J. E. McGrath, Chem. Rev. 2004, 104,<br />

4587.<br />

[30] M. J. Han, J. H. Park, J. Y. Lee, J. Y. Jho, Macromol. Rapid Comm. 2006, 27, 219.<br />

[31] S. Sareh, A. T. Conn, J. M. Rossiter, I. Ieropoulos, P. Walters, Proc. SPIE 2010, 7642,<br />

76421S.<br />

[32] P. Arena, C. Bonomo, L. Fortuna, M. Frasca, S. Graziani, IEEE Trans. Systems, Man, and<br />

Cybernetics B: Cybernetics 2006, 36, 1044.<br />

328


Abstract<br />

APPENDIX V<br />

Block Copolymer Self-Organization vs. Interfacial Modificationin<br />

Bilayered Thin-Film Laminates*<br />

Arif O. Gozen, Jiajia Zhou, Kristen E. Roskov, An-Chang Shi,<br />

Jan Genzer and Richard J. Spontak<br />

Block copolymers remain one of the most extensively studied and utilized classes of<br />

macromolecules due to their extraordinary abilities to (i) self-assemble spontaneously into a<br />

wide variety of soft nanostructures and (ii) reduce the interfacial tension between, and thus<br />

compatibilize, immiscible polymer pairs. In bilayered thin-film laminates of immiscible<br />

homopolymers, block copolymers are similarly envisaged to stabilize such laminates.<br />

Contrary to intuition, we demonstrate that highly asymmetric block copolymers can<br />

conversely destabilize a laminate, as discerned from macroscopic dewetting behavior, due to<br />

dynamic competition between copolymer self-organization outside and enrichment at the<br />

bilayer interface. The mechanism of this counterintuitive destabilization is interrogated<br />

through complementary analysis of laminates containing mixtures of stabilizing/destabilizing<br />

*This chapter has been published in its entirety<br />

AO Gozen, J Zhou, KE Roskov, AC Shi, J Genzer, RJ Spontak. Gozen, Arif O. "Block copolymer<br />

self-organization vs. interfacial modification in bilayered thin-film laminates." Soft matter 2011, 7,<br />

3268.<br />

329


diblock copolymers and time-dependent Ginzburg-Landau computer simulations. This<br />

combination of experiments and simulations reveal a systematic progression of<br />

supramolecular-level events that establish the relative importance of molecular aggregation<br />

and lateral interfacial structuring in a highly nonequilibrium environment.<br />

Introduction<br />

One of the most important technological challenges in the development of inexpensive<br />

polymeric materials with tailored properties is compatibilization of two or more immiscible<br />

homopolymers. 1 The demand for such polymeric systems continues to grow as the need for<br />

lightweight, processable and mechanically robust materials increases in response to efforts<br />

aimed at conserving natural resources and reducing energy consumption. 2 Most polymer<br />

pairs are inherently immiscible due to unfavorable thermodynamics: 3 the enthalpy of mixing<br />

is typically endothermic and the entropy of mixing is often negligibly small.<br />

Compatibilization requires a reduction in interfacial tension, which, in turn, is achieved<br />

through a variety of interfacial modification strategies. 4 One effective route is reactive<br />

blending, 5,6 which relies on the chemical coupling of dissimilar macromolecules at the<br />

polymer/polymer interface. The result is the in-situ development of block (or graft)<br />

copolymer molecules, which, when localized at the interface, serve to lower the interfacial<br />

tension (thereby reducing the size scale of phase separation), improve fracture toughness 7<br />

and restrict droplet coalescence. 8 While this approach is broadly applicable to a wide range<br />

of polymers and ensures that copolymer molecules reside at the interface where they are<br />

330


needed, it can suffer from undesirable side reactions, as well as low yields since the reactive<br />

macromolecules must meet at the interface for the reaction to proceed.<br />

An alternative to reactive blending relies on the physical addition of a premade block<br />

copolymer to an incompatible polymer blend. 9 In this case, the copolymer molecules must be<br />

allowed to diffuse to the developing polymer/polymer interface. While this can be largely<br />

accomplished through the use of plasticizing agents or high shear fields, a fraction of the<br />

copolymer population remains inevitably in one or both of the homopolymer phases. In this<br />

case, the copolymer molecules in the bulk seek to reduce unfavorable contacts with the<br />

surrounding matrix by self-assembling into nanostructured domains that protect their<br />

incompatible elements. Alone, diblock copolymers can organize spontaneously into periodic<br />

nanostructures ranging from spheres or cylinders of the minority component (on a body-<br />

/face-centered cubic or hexagonal lattice, respectively) in a matrix of the majority component<br />

to bicontinuous channels or lamellae. 10 In the presence of a solvent or homopolymer,<br />

aperiodic bicontinuous 11,12 and network 13 morphologies may also develop. Because of the<br />

propensity for block copolymers to self-organize, melt blending of incompatible<br />

homopolymers must be conducted in such fashion to keep the population of copolymer<br />

molecules remaining in one or both homopolymers relatively low; otherwise, trapped<br />

copolymer molecules form aggregates that resemble micelles, which prevent the interface<br />

from attaining its maximum strength and hinder compatibilization.<br />

Similar phenomena likewise occur during the copolymer-induced stabilization of<br />

bilayered thin-film laminates composed of two molecularly thin homopolymer films. Such<br />

331


laminates are important to the development of advanced protective coatings, 14 solar cells 15<br />

and waveguide assemblies, 16 in which case a detailed understanding of the molecular-level<br />

processes governing stabilization is required. In addition, the planar arrangement of such<br />

laminates provides a convenient test platform for the exploration of material and/or design<br />

variations in systematic fashion 17 while avoiding changes in interfacial curvature that would<br />

occur in bulk systems due to compatibilization. Without an added copolymer, a laminate may<br />

rupture either in a single layer or in both layers, 18,19 forming circular holes or other complex<br />

dewetting patterns, 20 when heated above their glass transition temperatures. Here, we only<br />

consider the cases where (i) the melt viscosity of the substrate layer is much larger than that<br />

of the top layer so that the substrate may be considered solid-like relative to the top layer; 21<br />

and (ii) destabilization proceeds by the nucleation and growth of rimmed holes that<br />

eventually impinge. 18 In the absence of interfacial slip, 22 the hole diameter (D) varies linearly<br />

with time (t), and the hole growth (dewetting) rate (dD/dt) depends on the ratio of the<br />

dewetting force to the friction caused by viscous dissipation. The magnitude of dD/dt affords<br />

a relative measure of interfacial stability and can, along with the mechanism of dewetting, be<br />

controlled by varying material parameters such as the thickness 23 or molecular weight 24 of<br />

the top layer (cf. Fig. A5.1), as well as adding a species that modifies the nature of the<br />

interface. 24,25<br />

We prepared thin-film laminates from two polystyrene (PS) homopolymers with number-<br />

average molecular weight ( M n) values of 30 and 50 kDa (PS30 and PS50, respectively), in<br />

conjunction with a poly(methyl methacrylate) homopolymer with<br />

��<br />

332<br />

��<br />

M n = 243 kDa


(PMMA243), all from Pressure Chemical, Inc. (Pittsburgh, PA). To the PS homopolymers,<br />

we added poly(styrene-b-methyl methacrylate) diblock copolymers (Polymer Source, Inc.,<br />

Dorval, Quebec, Canada) varying in molecular symmetry: S10M50, S50M54 and S50M10,<br />

where each numerical designation denotes the block molecular weight (in kDa). These<br />

materials, as well as solvent-grade toluene (Sigma-Aldrich, St. Louis, MO), were used as-<br />

received. Each PMMA243 substrate measuring 50±2 nm thick from ellipsometry was spun-<br />

cast at a speed of 2000 rpm onto a silicon wafer from a 1.35 wt% solution in toluene.<br />

Similarly, 1.55 wt% solutions of PS30 and PS50 in toluene were spun-cast at the same speed<br />

with and without added copolymers onto glass. Each film measured 60±2 nm thick and was<br />

floated off on deionized water and then deposited on a PMMA243 substrate to form a<br />

bilayered laminate. All laminates were dried for 24 h at ambient temperature and<br />

subsequently annealed at 180°C under nitrogen. Dewetting kinetics were monitored in<br />

reflection mode with an Olympus BX60 optical microscope equipped with a Mettler heating<br />

stage and a computer-interfaced CCD camera. Transmission electron microscopy (TEM;<br />

Hitachi HF2000, 200 kV) was performed on laminates prepared on silica-coated grids, and<br />

atomic force microscopy (AFM; Park Systems XE-100) was conducted in non-contact mode<br />

on specimens before and after selective removal of the top PS30 layer in 1-chloropentane. 26<br />

Dewetting rates are presented as a function of copolymer concentration for laminates<br />

composed of PS30 (Fig. A5.1A) and PS50 (Fig. A5.1B) top layers on PMMA243. In the<br />

absence of copolymer, measured values of dD/dt are 310±17 and 145±4 μm/h, respectively,<br />

confirming that a larger melt viscosity due to increased molecular weight of the top layer<br />

333


educes dD/dt and thus improves stability. 24 Similarly, these homopolymer dewetting rates<br />

can be used to distinguish copolymer-induced stabilization (or destabilization) by discerning<br />

whether dD/dt is lower (or higher) than that of the unmodified PS/PMMA243 interface. The<br />

dewetting rates with the addition of the asymmetric S50M10 and nearly symmetric S50M54<br />

copolymers (Fig. 1A) verify that both copolymers tend to promote stabilization, with<br />

S50M10 being more effective than S50M54. This comparison reveals that, with their long<br />

styrenic block and short methacrylic block, S50M10 molecules are less likely to form<br />

aggregates in PS30 and therefore diffuse to the interface where they physically intercalate the<br />

two homopolymers and reduce interfacial tension. An increase in copolymer concentration<br />

further improves the stability of PS30 due to a larger population of copolymer molecules<br />

available for interfacial modification. At very low concentrations, however, the S50M54<br />

copolymer is found to induce destabilization due to the presence of the longer, more PS-<br />

incompatible methacrylic block. In stark contrast to the general behavior of these<br />

copolymers, incorporation of the S10M50 copolymer fully destabilizes the PS30 layer over<br />

the entire copolymer concentration range examined (with no discernible concentration<br />

dependence). The molecular-level mechanism responsible for this unexpected and<br />

counterintuitive result is described and discussed below. Similar trends are evident in Fig. 1B<br />

for laminateswith PS50. Note, however, that in this case the dewetting rates measured for<br />

laminates with the S10M50 copolymer decrease with increasing concentration and approach<br />

that of the neat PS50.<br />

334


Comparison of the dewetting rates achieved by adding the S50M10 and S10M50 block<br />

copolymers (with identical molecular weights) in Fig. A5.1 indicates that S50M10 brings<br />

about stabilization, whereas S10M50 enhances destabilization. The mechanism by which the<br />

S50M10 copolymer improves the compatibility of the immiscible interface has already been<br />

discussed, but the behavior of the S10M50 copolymer is more complex, as illustrated in Figs.<br />

A5.1C-E. We hypothesize that the incompatibility between either PS30 or PS50 and the<br />

S10M50 molecules, each possessing a short styrenic block and a relatively long methacrylic<br />

block, is sufficiently high to induce the spontaneous formation of aggregates with<br />

methacrylic cores and styrenic shells, as evidenced by the inset of Fig. A5.1B. These<br />

aggregates measure 30±6 nm in diameter and resemble crew-cut micelles 27 (depicted in Fig.<br />

A5.1C). Because they are far from equilibrium, the copolymer molecules are likewise<br />

capable of adopting more complex shapes (e.g., vesicles or toroids). As the system evolves,<br />

aggregates and individual copolymer molecules migrate (at different rates), and ultimately<br />

fuse, to the interface, as portrayed in Fig. A5.1D. Existence of partially fused, as well as<br />

intact, aggregates along the interface causes increases in interfacial roughness and, hence,<br />

area, which promote an increase in free energy and destabilization of the top layer. In this<br />

case, the in-plane distribution of copolymer aggregates and molecules is not uniform.<br />

Eventual dissolution of aggregates into brush patches (cf. Fig. A5.1E) is expectedly related to<br />

the incompatibility between the styrenic matrix and the methacrylic block, which is greater in<br />

the PS50 than in the PS30 laminates. This consideration explains why dD/dt is (within<br />

335


experimental uncertainty) independent of copolymer concentration in Fig. A5.1A, but<br />

decreases noticeably with increasing copolymer concentration in Fig. A5.1B.<br />

Evidence supporting our proposed mechanism can be attained from time-dependent<br />

Ginzburg-Landau computer simulations, which were performed with the assumption that the<br />

PMMA243 substrate layer can be treated as a PMMA-attractive surface to simplify and<br />

accelerate the calculations. A free energy functional proposed 28 previously for AB diblock<br />

copolymer/homopolymer blends is incorporated into the Cahn-Hilliard equation to model<br />

short-time system dynamics. The spatiotemporal behavior of two asymmetric copolymers<br />

configured to emulate the S10M50 and S50M10 molecules, one with 17% A (A17B83) and<br />

the other with 83% A (A83B17), is considered in a homopolymer A matrix in terms of (i) an<br />

order parameter (�) that reflects the local copolymer concentration and (ii) local height<br />

variations that provide a measure of roughness and, hence, structuring. The time-dependent<br />

variation of � in the z direction, where z is normal to the A/B interface (Fig. A5.3A), shows<br />

that the concentrations of both copolymer molecules along the interface (z = 0) increase as<br />

the system evolves. A striking difference between the two species is that the A17B83<br />

molecules extend from the interface as organized aggregates (evidenced by the fluctuations<br />

in �), whereas the A83B17 molecules do not. A 2D image of a lateral simulation near the<br />

interface for a laminate with A17B83 molecules is provided in the inset of Fig. A5.2A, and<br />

reveals the existence of a complex copolymer morphology loosely reminiscent of the<br />

nanostructure in Fig. 1B. In contrast, a corresponding image of the A83B17 molecules<br />

displays significantly less lateral structuring. Root-measured (rms) roughness values<br />

336


extracted from such 2D simulations aregiven in Fig. A5.2B and confirm that the A17B83<br />

molecules are more organized, especially near the interface, than the A83B17 molecules,<br />

which is consistent with our proposed mechanism.<br />

Experimental AFM measurements of the interfacial roughness of the PS30/S10M50<br />

laminate after selective removal of the PS30 layer are included for comparison in the inset of<br />

Fig. A5.2B and indicate that, at short times, the roughness discerned from both dry (i.e.,<br />

dewetted) and wet (i.e., not dewetted) regions on the PMMA243 surface increases as<br />

dewetting proceeds, in agreement with simulation results (at different time scales). This<br />

increase in roughness is attributed to the attachment and partial fusion of copolymer<br />

aggregates along the interface. At longer times, the roughness decreases as copolymer<br />

aggregates meld into the PMMA243 substrate. This process is observed and expected to be<br />

faster (and more complete) for dry regions exposed to surface tension than for wet regions<br />

subjected to lower interfacial tension. Existence of interfacial copolymer structuring due to at<br />

least partially fused aggregates is verified by the TEM images presented in Figs. A5.3A and<br />

A5.3B for laminates containing 0.15 and 0.75 wt% S10M50, respectively, after 6 min at<br />

180°C. The dark features on the dry PMMA243 surface distinguish stained styrenic moieties<br />

and serve to indicate copolymer-rich interfacial regions. At the low copolymer concentration<br />

(Fig. A5.3A), discrete features possess diameters up to 35 nm, which is consistent with the<br />

size of copolymer aggregates measuring ≈4Rg, where Rg denotes the copolymer gyration<br />

radius (≈7 nm). At the higher concentration (Fig. A5.3B), these features are irregularly<br />

337


shaped and possess a broad size distribution extending up to several hundred nanometers<br />

across.<br />

According to experimental observations and simulation results, self-assembly of the<br />

S10M50 copolymer molecules occurs rapidly, resulting in the formation of micelle-like<br />

aggregates that migrate to and roughen the polymer/polymer interface, consequently<br />

destabilizing the top PS layer. In contrast, the mirrored S50M10 copolymer behaves in<br />

largely opposite fashion: individual copolymer molecules diffuse to and meld with the<br />

interface, where they stabilize the laminate. To discern the relative importance of these<br />

competitive molecular-level mechanisms, we have prepared laminates containing mixtures of<br />

these two copolymers and measured the dewetting rates, which are presented in Fig. A5.4. At<br />

1 and 2 wt% S50M10, the destabilization mechanism dominates. Here, the population of<br />

S50M10 molecules is insufficient to modify the polymer/polymer interface, whereas the<br />

remaining (and more numerous) S10M50 chains favor self-assembly over interfacial<br />

modification. Between 3 and 5 wt% S50M10, however, destabilization at low concentrations<br />

precedes stabilization. Stabilization is achieved to different extents by having as little as 10<br />

wt% S50M10 in the S10M50/S50M10 mixture. As seen in Fig. A5.4, using block copolymer<br />

mixtures rather than single copolymers to tune stabilizing/ compatibilizing efficacy provides<br />

an unexplored route to achieving property control from the ground up. Such control must<br />

consider the complex interplay between block copolymer self-assembly and interfacial<br />

modification under highly nonequilibrium conditions. Our results using a planar test<br />

338


configuration elucidate a molecular-level mechanism responsible for this interplay, which is<br />

of critical importance to the contemporary development of tailored polymeric materials.<br />

Acknowledgments<br />

Graduate fellowships for A.O.G. and K.E.R. were provided by Dade-Behring, Inc. and<br />

the National Science Foundation, respectively. J.Z. and A.C.S. were supported by the Natural<br />

Science and Engineering Council of Canada. The computer simulations were made possible<br />

by the facilities of the Shared Hierarchical Academic Research Computing Network<br />

(SHARCNET: www.sharcnet.ca) and Compute/Calcul Canada.<br />

339


Figures<br />

Figure A5.1. Dewetting rates (dD/dt) presented as a function of copolymer concentration for<br />

laminates with (A) PS30 and (B) PS50 as the top PS layer (labeled in A). Diblock<br />

copolymers defined in the text and added to the top layer are likewise color-coded and<br />

labeled. Solid lines connect the data, and the dashed line delineates top-layer stabilization<br />

from destabilization. Error bars denote one standard deviation in the data. The TEM image in<br />

(B) shows the ill-defined and faint S10M50 nanostructure that develops upon spin-casting a<br />

PS50/S10M50 film with 0.75 wt% S10M50 onto glass, followed by floated transfer onto a<br />

PMMA243 substrate layer spun-cast on a silica-coated grid. In this image, styrenic units are<br />

stained with the vapor of RuO4(aq) so that unstained methacrylic moieties appear light.<br />

Examples of aggregates resembling micelles are circled, whereas more complex shapes are<br />

identified by arrowheads. Included are schematic illustrations of the mechanism by which<br />

asymmetric S10M50 block copolymers destabilize a bilayered laminate: (C) copolymer<br />

molecules self-organize into aggregates (shown here), as well as more complex<br />

nanostructural elements (cf. the inset in B) upon initial casting; (D) copolymer aggregates<br />

and chains in the melt diffuse to the polymer/polymer interface where they adsorb and<br />

eventually undergo fusion, promoting an increase in interfacial roughness; and (E) additional<br />

aggregates form (depending on the available copolymer reservoir) and continue to migrate to<br />

and meld with the interface to form copolymer brushes.<br />

340


Figure A5.2. In (A), the concentration-based order parameter (�) presented as a function of<br />

distance from the A/B polymer interface (at z = 0) at different simulation times (τ, labeled)<br />

for A17B83 and A83B17 copolymer molecules (labeled and defined in the text) at a<br />

copolymer concentration of 1.0 wt%. A pair of 2D lateral simulation images near the A/B<br />

interface at τ = 40 is displayed for both copolymers (labeled) in the inset of (A). Values of the<br />

rms roughness (in lattice units, l.u.) extracted from simulation images such as those provided<br />

in (A) are provided for the A17B83 ( ) and A83B17 ( ) copolymers as a function of τ, in<br />

(B). Included in the inset of (B) are experimental rms roughness values measured by AFM of<br />

wet (not dewetted, ) and dry (dewetted, ) interfacial regions of a laminate after selective<br />

removal of the PS30/S10M50 top layer. The solid lines connect the data.<br />

341


Figure A5.3. TEM images acquired from dry regions of annealed laminates with<br />

PS50/S10M50 top layers on PMMA243 at two S10M50 concentrations (in wt%): (A) 0.15<br />

and (B) 0.75. Styrene-containing features remaining on the PMMA243 substrate after<br />

dewetting appear electron-opaque (dark).<br />

342


Figure A5.4. Dewetting rates presented as a function of total copolymer concentration for<br />

PS50/PMMA243 bilayered laminates with and without mixtures of the asymmetric S50M10<br />

and S10M50 block copolymers at different mixture compositions (color-coded, labeled and<br />

expressed in w/w S10M50/S50M10). Solid lines connect the data, and the error bars denote<br />

one standard deviation in the data.<br />

343


References<br />

1. U. Sundararaj, C. W. Macosko, Macromolecules 28, 2647 (1995).<br />

2. P. Ball, Made to Measure: New Materials for the 21 st Century (Princeton University Press,<br />

Princeton, NJ, 1997).<br />

3. P. J. Flory, Principles of Polymer Chemistry (Cornell University Press, Ithaca, NY, 1953).<br />

4. C. E. Koning, M. van Duin, C. Pagnoulle, R. Jérôme, Prog. Polym. Sci. 23, 707 (1998).<br />

5. C. A. Orr et al., Polymer 42, 8171 (2001).<br />

6. H. Pernot, M. Baumert, F. Court, L. Leibler, Nat. Mater. 1, 54 (2002).<br />

7. C. Creton, E. J. Kramer, C.-Y. Hui, H. R. Brown, Macromolecules 25, 3075 (1992).<br />

8. S. Lyu, T. D. Jones, F. S. Bates, C. W. Macosko, Macromolecules 35, 7845 (2002).<br />

9. A. V. Ruzette, L. Leibler, Nat. Mater. 4, 19 (2005).<br />

10. I. W. Hamley, The Physics of Block Copolymers (Oxford Univ. Press, NY, 1998).<br />

11. H. Jinnai et al., Adv. Mater., 14, 1615 (2002).<br />

12. P. Falus, H. Xiang, M. A. Borthwick, T. P. Russell, S. G. J. Mochrie, Phys. Rev. Lett. 93,<br />

145701 (2004).<br />

13. S. Jain, F. S. Bates, Science 300, 460 (2003).<br />

14. C. K. Tan, D. J. Blackwood, Corros. Sci. 45, 545 (2003).<br />

15. E. J. W. Crossland et al., Nano Lett. 9, 2807 (2009).<br />

16. D. H. Kim et al., Adv. Mater. 17, 2442 (2005).<br />

17. S. Zhu et al., Nature 400, 49 (1999).<br />

18. G. Reiter, Phys. Rev. Lett. 68, 75 (1992).<br />

19. J. P. de Silva et al., Phys. Rev. Lett. 98, 267802 (2007).<br />

344


20. R. Xie, A. Karim, J. F. Douglas, C. C. Han, R. A. Weiss, Phys. Rev. Lett. 81, 1251 (1998).<br />

21. F. Brochard-Wyart, P. Martin, C. Redon, Langmuir 9, 3682 (1993).<br />

22. K. Jacobs, R. Seemann, G. Schatz, S. Herminghaus, Langmuir 14, 4961 (1998).<br />

23. R. Limary, P. F. Green, Macromolecules 32, 8167 (1999).<br />

24. B. Wei, J. Genzer, R. J. Spontak, Langmuir 20, 8659 (2004).<br />

25. B. Wei, P. G. Lam, J. Genzer, R. J. Spontak, Langmuir 22, 8642 (2006).<br />

26. S. E. Harton, J. Luning, H. Betz, H. Ade, Macromolecules 39, 7729 (2006).<br />

27. L. F. Zhang, A. Eisenberg, Science 268, 1728 (1995).<br />

28. T. Ohta, A. Ito, Phys. Rev. E 52, 5250 (1995).<br />

345

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!