You are on page 1of 12

CRC_2308_Ch019.

qxd 12/1/2005 3:46 PM Page 553

19 Nanofiber Technology
Frank K. Ko
Department of Materials Science and Engineering,
Drexel University, Philadelphia, Pennsylvania

CONTENTS

19.1 Introduction
19.2 The Electrospinning Process
19.3 Key Processing Parameters
19.4 Nanofiber Yarns and Fabrics Formation
19.5 Potential Applications of Electrospun Fibers
19.5.1 Nanofibers for Tissue Engineering Scaffolds
19.5.2 Nanofibers for Chemical/Bio Protective Membranes
19.5.3 Nanocomposite Fibers for Structural Applications
19.6 Summary and Conclusions
References

19.1 INTRODUCTION
Nanofiber technology is a branch of nanotechnology whose primary objective is to create materials
in the form of nanoscale fibers in order to achieve superior functions. The unique combination of
high specific surface area, flexibility, and superior directional strength makes such fibers a preferred
material form for many applications ranging from clothing to reinforcements for aerospace struc-
tures. Although the effect of fiber diameter on the performance and processibility of fibrous struc-
tures has long been recognized, the practical generation of fibers at the nanometer scale was not
realized until the rediscovery and popularization of the electrospinning technology by Professor
Darrell Reneker almost a decade ago [1]. The ability to create nanoscale fibers from a broad range
of polymeric materials in a relatively simple manner using the electrospinning process, coupled with
the rapid growth of nanotechnology in recent years have greatly accelerated the growth of nanofiber
technology. Although there are several alternative methods for generating fibers in a nanometer scale,
none matches the popularity of the electrospinning technology due largely to the simplicity of the
electrospinning process. Electrospinning can be carried out from polymer melt or solution. A major-
ity of the published work on electrospinning has been focused on solution-based electrospinning
rather than on melt electrospinning due to higher capital investment requirements and the difficulty
in producing submicron fibers by melt electrospinning. We will concentrate on solution-based elec-
trospinning in this chapter. Specifically, after an introduction to the processing principles the relative
importance of the various processing parameters in solution electrospinning is discussed. The struc-
ture and properties of the fibers produced by the electrospinning process are then examined.
Recognizing the enormous increase in specific fiber surface, bioactivity, electroactivity, and enhance-
ment of mechanical properties, numerous applications have been identified, including filtration, bio-
medical, energy storage, electronics, and multifunctional structural composites.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 554

554 Nanomaterials Handbook

19.2 THE ELECTROSPINNING PROCESS


Electrostatic generation of ultrafine fibers or ‘‘electrospinning’’ has been known since the 1930s [2].
The rediscovery of this technology has stimulated numerous applications including high-performance
filters [3] and as scaffolds in tissue engineering [4],which utilize the unique characteristics of surface
area as high as 103 m2/g provided by nanofibers. In this nonmechanical, electrostatic technique, a high
electric field is generated between a polymer fluid contained in a glass syringe with a capillary tip and
a metallic collection target. When the voltage reaches a critical value, the electric field strength over-
comes the surface tension of the deformed droplet of the suspended polymer solution formed on the
tip of the syringe, and a jet is produced. The electrically charged jet undergoes a series of electrically
induced bending instabilities during its passage to the fiber collection screen or drum that results in
the hyperstretching of the jet. This stretching process is accompanied by the rapid evaporation of the
solvent molecules that reduces the diameter of the jet, in a cone-shaped volume called the ‘‘envelope
cone.’’ The dry fibers are accumulated on the surface of the collection screen resulting in a nonwoven
random fiber mesh of nano- to micron diameter fibers. The process can be adjusted to control the fiber
diameter by varying the electric field strength and polymer solution concentration. By proper control
of the electrodes, aligned fibers can also be produced. A schematic drawing of the electrospinning
process and the random and aligned nanofibers are shown in Figure 19.1.

Metering (b)
pump

Syringe
Taylor cone (Polymer solution) 1 µm bar
stability region
instability region +
nanofibers
V
High voltage
Target/collection power supply
plate

(a) (c)

FIGURE 19.1 (a) Schematic drawing of the electrospinning process. (b) (top) random; (c) (bottom) aligned.

(a) (b)

FIGURE 19.2 (a) The first commercially available nanofiber electrospinning unit (NEU) produced by the
Kato-Tech Co. (b) Close-up of the NEU system including a metering pump, a syringe containing the spinning
dope, the syringe needle is connected to a power supply (not shown), and a fiber collection drum connecting
to the electrical ground.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 555

Nanofiber Technology 555

Figure 19.2 shows a commercial nanofiber electrospinning unit known as NEU produced by the
Kato-Tech Co. The NEU system consists of a metering pump, which controls the volume flow rate
of the spinning dope in a syringe. An electrical potential is applied by a power supply between the
steel syringe needle and the fiber collection ground in the form of a metallic screen or a drum. The
temperature and humidity of the spinning chamber can be controlled.

19.3 KEY PROCESSING PARAMETERS


A key objective in electrospinning is to generate fibers of nanometer diameter consistently and
reproducibly. Considerable effort has been devoted to understand the parameters affecting the
spinnability and more specifically the diameter of the fibers resulting from the electrospinning
process. Many processing parameters have been identified, which influence the spinnability and the
physical properties of nanofibers. These parameters include electric field strength, polymer con-
centration, spinning distance, and polymer viscosity.
According to Rutledge et al. [5], the diameter of electrospun fibers is governed by the following
equation:
1/3
Q2
冤 冥
2
d ⫽ γε ᎏ ᎏᎏ
I 2 π (2ln χ ⫺3)

where d is the fiber diameter, γ the surface tension, ε the dielectric constant, Q the flow rate, I the
current carried by the fiber, and χ the ratio of the initial jet length to the nozzle diameter.
One can control the fiber diameter by adjusting the flow rate, the conductivity of the spinning
line, and the spinneret diameter. Fiber diameter can be minimized by increasing the current carry-
ing capability of the fiber through the introduction of a conductive filler such as carbon black, car-
bon nanotube, metallic atoms, or mixing with an inherently conductive polymer. For example, an
increase of current carrying capability, I, of the fiber by 32 times will bring about 10-fold decrease
in fiber diameter. Alternatively, if the current I is kept constant, one can bring about a 10-fold
decrease in the fiber diameter by reducing the flow rate by 32 times. Reducing the spinneret nozzle
diameter can also bring about a reduction in fiber diameter. When the value of χ is increased from
10 to 1000, the diameter of the fiber will decrease by approximately 2 times.
Experimental evidence has shown that the diameter of the electrospun fibers is influenced by
molecular conformation that is related to the molecular weight and the concentration of the poly-
mer in the spinning dope [6]. It was found that the diameter of fibers spun from dilute polymer solu-
tions can be expressed in terms of the Berry number B, a dimensionless parameter and a product of
intrinsic viscosity η, and polymer concentration, C. This relationship has been observed in a large
number of polymers. An example for this relationship is illustrated using polylactic acid (PLA). The
relationship between polymer concentration and fiber diameter of a PLA of different molecular
weight in chloroform are shown in Figure 19.3. It can be seen that fiber diameter increases as poly-
mer concentration increases. The rate of increase in fiber diameter is greater at higher molecular
weight. Accordingly, one can tailor the fiber diameter by proper selection of polymer molecular
weight and polymer concentrations.
Expressing fiber diameter as a function of B, as shown in Figure 19.4, a pattern emerges illus-
trating four regions of B vs. diameter relationships. At region (I), B⬍1, characterizing a very dilute
polymer solution with molecular chains barely touching each other. It is almost impossible to form
fibers by electrospinning of such solution, since there is not enough chain entanglement to form a
continuous fiber, and the effect of surface tension will make the extended conformation of a single
molecule unstable. As a result, only polymer droplets are formed. In region (II), 1⬍B⬍3, fiber
diameter increases slowly as B increases within the range ⬃100 to ⬃500 nm. In this region, the
degree of molecular entanglement is just sufficient for fiber formation. The coiled macromolecules
of the dissolved polymer are transformed by the elongational flow of the polymer jet into orientated

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 556

556 Nanomaterials Handbook

Average diameter vs. Concentration (Mw=300,000)


4000
R 2 = 0.8478
3500 y = 895.29 ×−1648.5

3000
Average diameter (nm)
2500

2000

1500

1000

500

0
0 1 2 3 4 5 6
Concentration (g /dL)

Average diameter vs. Concentration (Mw=200,000)


3500
y = 330.3 × −1018.4
3000 R 2 = 0.9005
Average diameter (nm)

2500

2000

1500

1000

500

0
0 5 10 15
Concentration (g/dL)

FIGURE 19.3 Relationship of fiber diameter and concentrations of polymer with different molecular weights.

molecular assemblies with some level of inter- and intramolecular entanglement. These entangled
networks persist as the fiber solidifies. In this region, some bead formations are observed as a result
of polymer relaxation and surface tension effect. In region (III), 3⬍B⬍4, fiber diameter increases
rapidly with B in the range from ⬃1700 to ⬃2800 nm. In this region, the molecular chain entan-
glement becomes more intensive, contributing to an increase in polymer viscosity. Because of the
intense level of molecular entanglement, it requires a higher level of electric field strength for fiber
formation by electrospinning. In region (IV), B⬎4, the fiber diameter is less dependent on B. With
a high degree of inter- and intramolecular chain entanglement, other processing parameters such as
electric field strength and spinning distance become dominant factors affecting fiber diameter. A
schematic illustration of the four Berry regions is shown in Table 19.1, showing the corresponding
relationship between fiber morphology and Berry numbers.
More recently, we further demonstrated, as shown in Figure 19.5, that the diameter, d, of elec-
trospun fibers is related to the Berry number B for three amorphous polymers with different molec-

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 557

Nanofiber Technology 557

Average diameter vs. Berry number


3000

2500

Average diameter (nm)


2000

1500 I II III IV

1000

500

0
0 1 2 3 4 5
Berry number

FIGURE 19.4 The relationship between Berry number and fiber diameter.

TABLE 19.1
Schematic of Polymer Chain Conformation and Fiber Morphology Corresponding to Four
Regions of Berry Number
Region I Region II Region III Region IV
Berry Number B<1 1<B<2.7 2.7B<3.6 B>3.6

Polymer chain conformation in


solution

Fiber morphology

Average fiber diameter (Only droplets ⬃100–500 nm 1700–2800 nm ⬃2500–3000 nm


formed)

ular weights. It was shown that the relationship of fiber diameter to Berry number can be expressed
by d ⫽ a Bc, where a and c are experimental coefficients. The molecular weight of the three poly-
mers polystyrene, polybutadiene, and SBS are 48,000, 60,000, and 100,000 dalton, respectively. It
is of interest to note that the slope of the diameter vs. Berry number curves are approximately the
same, with a value of 5, reflecting the degree of crystallinity of the polymer. The coefficient a· is
thought to be related to the molecular weight, radius of gyration of molecular chains, and entan-
glement of the molecular chains. From this relationship, the polymer concentration necessary to
produce nanofiber can be predicted.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 558

558 Nanomaterials Handbook

100
Dia. = B 5

Fiber diameter (micron meter)


10 SBS(TR2000)
PB(RB830)
Polystyrene
1 PB/VGCF(1phr)

0.1
1 10
B

FIGURE 19.5 The relationship between fiber diameter and the Berry number.

19.4 NANOFIBER YARNS AND FABRICS FORMATION


In order to translate nanoeffect to the macroscopic structural level, the nanofibers must be organ-
ized into linear, planar, and 3-D assemblies. The formation of nanofiber assemblies provides a
means to connect nanofiber effect to macrostructure performance. Electrospun fiber structures can
be assembled by direct fiber-to-fabric formation to create a nonwoven assembly and by the creation
of a linear assembly or a yarn from which a fabric can be woven, knitted, or braided. Linear fiber
assemblies can be aligned mechanically or by electrostatic field control. Alternatively, a self-assem-
bled continuous yarn can be formed during electrospinning by proper design of the ground elec-
trode. Self-assembled yarn can be produced in continuous length with appropriate control of
electrospinning parameters and conditions. The fibers are allowed to accumulate until a tree-like
structure is formed. Once a sufficient length of yarn is formed, the accumulated fibers attach them-
selves to the branches and continue to build up. A device such as a rotating drum can be used to
spool up the self-assembled yarn in continuous length as shown in Figure 19.6. This method pro-
duces partially aligned nanofibers yarn bundles of continuous length.

19.5 POTENTIAL APPLICATIONS OF ELECTROSPUN FIBERS


The Donaldson Company was the first to realize the commercial value of electrospinning taking
advantage of the enormous availability of the specific surface of electrospun fibers and the ultrafine
nature of the fibers. They introduced the Ultra-Web® cartridge filter for industrial dust collection in
1981 and more recently the Hollingsworth & Vose Company introduced the Nanoweb® for auto-
motive and truck filter applications. To date, industrial filter remains the primary commercial use of
electrospun fibers. The rediscovery and popularization of the electrospinning process in the 1990s
have changed the dynamics of the field of nanofiber technology. Electrospinning has been trans-
formed from an obscure technology to a word common in academia and one gradually being rec-
ognized by industry and government organizations. This was most evident in the Polymer Nanofiber
Symposium held in New York at the American Chemical Society (ACS) in September 2003,
wherein 75 papers and posters were presented [7]. On the basis of the papers presented in the ACS
symposium and the rapidly growing number of publications, one can categorize the potential
applications of electrospun fibers into (1) biomedical applications — tissue engineering scaffolds,
wound care, superabsorbent media, drug delivery carrier, etc. [8,9]; (2) electronic applications —
electronic packaging, sensors, wearable electronics, actuators, fuel cells, etc. [10,11]; (3) industrial
applications — filtration, structural toughening/reinforcement, chemical/bio-protection, etc.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 559

Nanofiber Technology 559

High voltage
V
power supply

Polymer
solution

Self-assembled
yarn

Rotary drum

FIGURE 19.6 Schematic illustration of the formation of self-assembled yarns by the electrospinning process.

[12,13]. It is of interest to note that these applications invariably made use of the large surface area
created by the fineness of the fibers. A few examples will be used herein to illustrate the potential
applications of nanofibers.

19.5.1 NANOFIBERS FOR TISSUE ENGINEERING SCAFFOLDS


It is well recognized, as articulated by Vacanti and Mikos [14], that the key challenges in tissue engi-
neering are the synthesis of new cell adhesion-specific materials and the development of fabrication
methods to produce reproducible three-dimensional synthetic or natural biodegradable polymer scaf-
folds with tailored properties. These properties include porosity, pore size distribution and connec-
tivity, mechanical properties for load-bearing applications, and rate of degradation. Of particular
interest in tissue engineering is the surface adhesion of the cells to the scaffold. Considering the
importance of surfaces for cell adhesion and migration, experiments were carried out in our labora-
tory using osteoblasts isolated from neonatal rat calvarias and grown to confluence in Ham’s F-12
medium (GIBCO), supplemented with 12% sigma fetal bovine on PLAGA-sintered spheres, 3-D-
braided 20 µm filament bundles, and nanofibrils. Four matrices were fabricated for the cell culture
experiments. These matrices include (1) 150–300 µm PLAGA-sintered spheres, (2) unidirectional
bundles of 20 µm filaments, (3) 3-D-braided structure consisting of 20 bundles of 20 µm filaments,
and (4) nonwoven consisting of nanofibrils. The cell proliferation, as shown in Figure 19.7, is
expressed in terms of the amount of [3H]-thymidine uptake as a function of time. It can be seen that
there is a consistent increase in cell population with time for all the matrices. However, the nanofi-
brous structures demonstrated the most proliferate cell growth, whereas the cell growth in the struc-
tures consisting large diameter fibers and spheres were the least proliferate. Taking advantage of this
cell-friendly characteristic of nanofiber scaffolds, nanofibers are being evaluated for the regeneration
of skin, vascular grafts, ligament, cartilage, bone, and many other tissues and organs.

19.5.2 NANOFIBERS FOR CHEMICAL/BIO PROTECTIVE MEMBRANES


Nanofiber assemblies are excellent structure barrier systems for dust filtration [12] and for the pre-
vention of the penetration of chemical and biological agents [13]. Considering the various barrier
concepts shown in Figure 19.8, it is concluded that neither the impermeable nor the permeable

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 560

560 Nanomaterials Handbook

Cell proliferation

3-D braid A
Thymidine uptake (µCi)

0.007
0.006 Microsphere B
0.005 Nano nonwoven C
0.004 Control D C
0.003
0.002 D
B
0.001 A
0
1 3 7 10 11
Time (days)

FIGURE 19.7 Directed cell growth (top) and cell proliferation (bottom) behavior in PLA nanofiber scaffold.

barrier systems are suitable because of the unacceptable level of heat generation in the imperme-
able barriers and the porous nature of the permeable systems. The semi-permeable barrier provides
incremental improvement but the addition of the sorptive layer may serve as a reservoir for bacte-
ria growth as well as adding bulk to the garment. Accordingly, the current design favors a system
that is carbon-free with selective permeable capability and which has the multiple functions of
repelling CB agents while allowing the body moisture vapor to escape. As an example, we illustrate
the feasibility of using the electrospinning process to construct a selective permeable membrane
using the polymers formulated and provided by Utility Development Cooperation (UDC).
An example of an solvent-based electrospun polymer is shown in Figure 19.9 showing a three
dimensional interconnected fiber network. The morphology of the polymer was found to be con-
trollable by varying polymer concentration.
The electrospun membranes were tested at UDC for moisture vapor transmission by using the
Thwing Albert permeation cups. The testing was performed at 110°F to simulate hot weather condi-
tions. The moisture vapor transmission results of dip-coated nylon fabric with the same polymer as the
electrospun polymer formulation were compared to uncoated nylon fabric. These dip-coated fabrics
were about 6 mil thick compared to electrospun membrane at 10 mil thick. As shown in Table 19.2,
moisture vapor transmission through electrospun membrane was equivalent to uncoated nylon fabric
and there was no detectable moisture vapor transmission through the dip-coated nylon fabric. The test
duration was 5 h. Considering that the water vapor permeability of the state-of-the-art protective
membranes are of the order of 0.1 to 0.25 g/cm2, these results indicated that the electrospun mem-
branes meet the requirements for wearing comfort over an extended period. This favorable compari-
son with the performance of the state-of-the-art membranes appears to be consistent with the findings

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd
Nanofiber Technology

12/1/2005
3:46 PM
Air Vapor Air Vapor Air Air
Vapor Vapor
Liquids Aerosols Liquids Aerosols Liquids Aerosols Liquids Aerosols

Page 561
Permeable Semipermeable Selectively permeable Impermeable

Sorptive material Sorptive material


Moisture Moisture Moisture
Moisture
vapor vapor vapor
vapor

Skin Skin Skin Skin

This is Natick's
current R&D focus

Selectively semipermeable barrier construction, Source: Wilusz, 1998.

FIGURE 19.8 Concepts of barrier systems for chemical/biological agent protection.

Copyright 2006 by Taylor & Francis Group, LLC

561
CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 562

562 Nanomaterials Handbook

FIGURE 19.9 Electrospun UDC843A polymer showing an interconnected fiber network.

TABLE 19.2
Water Vapor Loss of Various Membranes (g/cm2)
Uncoated Nylon Electrospun Membrane Dip Coated Nylon Cloth
on Nylon Cloth
0.402 0.348 0.055
0.323 0.105 0.012
0.478 0.235 0.035

of Gibson et al. [15], who showed that the water-transport properties of electrospun nonwovens are
superior to that of the state-of-the-art commercial membranes.
Chemical agent testing was performed at Geomet Technology using HD chemical agents for
these tests in accordance with MIL-STD-282 static diffusion test. As indicated in the report from
Geomet, the UDC-formulated product has more than 24 h of HD chemical agent resistance exceed-
ing the performance requirement for state-of-the-art chemical/bio protective membranes.

19.5.3 NANOCOMPOSITE FIBERS FOR STRUCTURAL APPLICATIONS


Carbon nanotubes (CNTs) [16] are seamless graphene tubule structures with nanometer-size diame-
ters and high aspect ratios. This new class of one-dimensional material is shown to have exceptional
mechanical, thermal, and novel electronic properties. The elastic moduli of the CNTs are in the range
of 1 to 5 TPa [17–19] and fracture strains of 6 to 30%—both parameters about an order of magnitude
better than those of commercial carbon fibers, which typically have 0.1 to 0.5 TPa elastic moduli and
0.1 to 2% fracture strains [20]. The factor of 10 enhancement in strength implies that, for the same
performance, replacing or augmenting commercial carbon fibers with CNTs will lead to significant
reduction in the volume and weight of the structural composites currently used in space applications.
Studies carried out in our laboratory have demonstrated that composite nanofibers containing 1
to 10 wt% single wall nanotube (SWNT) can be produced by electrospinning [21,22]. Fiber diam-
eters in the range of 40 to 300 nm have been obtained by varying the solution concentration and
electrospinning parameters. The alignment of SWNT in the nanofibers was confirmed by Raman
spectroscopy and TEM. Bulk mechanical properties of the green (as-spun) composite nanofibers
were characterized in the random mat form. With proper dispersion of SWNT and preparation of
the spinning dope, a twofold increase in strength and modulus is observed for the polyacrylonitrile
(PAN)/SWNT fiber assemblies containing 1 wt% SWNT. In addition, a continuous process for the
conversion of nanofibers into continuous yarn has also been successfully demonstrated. These con-
tinuous yarns provide an effective means to translate the properties of the SWNT to higher order
textile structures suitable for advanced composites and numerous other applications.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 563

Nanofiber Technology 563

35
Aligned
1 wt%
30 SWNT/PAN
fibers
25

Stress (MPa) 20 Random


1 wt%
SWNT/PAN
15 fibers

10

5 Random
PAN fibers
0
0 0.05 0.1 0.15 0.2
Strain

FIGURE 19.10 Tensile properties of electrospun pristine PAN and 1 wt% SWNT/PAN fibers in random and
aligned.

Figure 19.10 shows the stress–strain properties of electrospun pristine PAN and 1 wt%
SWNT/PAN nanofibers. As can be seen in the random fiber mats, a more than twofold increase in
tensile strength and elastic modulus was obtained with the addition of 1 wt% SWNT. For the
aligned 1 wt% SWNT/PAN, a 33% increase in tensile strength and almost double strain were
obtained as compared with the random composite nanofibers. However, the increase in elastic mod-
ulus is not significant. A large increase in the area under the stress–strain curve of SWNT/PAN indi-
cates the effectiveness of SWNT in toughening and strengthening the fibers. Fibril as well as SWNT
alignments play an important role in maximizing properties translation from the nanoscopic level
to higher order structure. Even without any posttreatment, remarkable improvement in properties
was achieved in the electrospun composite nanofibrils. It is anticipated that posttreatment such as
mechanical drawing further aligns the SWNT in the fiber axis. Drawing of the fibers also induces
molecular orientation resulting in much higher properties. Furthermore, heat treatment enhances
interaction between the reinforcement and matrix leading to better load transferring across rein-
forcement–matrix interface.
The stress–strain behavior of the SWNT-loaded PAN suggested toughening and strengthening
of the fiber with the incorporation of 1 wt % SWNT.

19.6 SUMMARY AND CONCLUSIONS


The combination of high-specific surface area, flexibility, and superior directional strength makes
fiber a preferred material form for many applications ranging from clothing to reinforcements for
aerospace structures. The availability of fibers in nanoscale will greatly expand the performance limit
of materials and thus create new marketing opportunities leading to the birth of a new nanofiber-
based industry. Of the processing methods capable of fabricating nanofibers, electrospinning has
received the most intensive attention due to its simplicity and the potential for generating continuous
nanofiber assemblies. With the enormous increase in available surface area per unit mass, nanofibers
provide a remarkable capacity for the attachment or release of functional groups, absorbed mole-
cules, ions, catalytic moieties, and nanometer scale particles of many kinds. As a result, these
nanofiber assemblies provide favorable conditions for cell adhesion and cell migration for bioactive
nanofibers thus creating new opportunities for nanofiber-based scaffolds for tissue engineering and
for ultra-sensitive biosensors. Other areas expected to be impacted by nanofiber-based technology

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch019.qxd 12/1/2005 3:46 PM Page 564

564 Nanomaterials Handbook

include drug delivery systems, wires, capacitors, transistors, and diodes for information technology,
systems for energy transport, conversion and storage, such as batteries and fuel cells, and structural
composites for aerospace structures. In order to capitalize on these new opportunities new process-
ing technology is needed, precise characterization tools for single nanofibers are required, and
nanoscale manipulation machines must be developed. These technologies must be supported by well-
verified engineering design tools and analytical methods. It is expected that nanofiber technology
will provide a critical link between nanoscale effect and macroscopic phenomena and become the
enabling pathway for the conversion of nanomaterials into practical large-scale applications.

REFERENCES
1. Reneker, D.H. and Chun, I., Nanotechnology, 7, 216, 1966.
2. Formhals, A., US Patent 1,975,504, October 2, 1934.
3. Donaldson Co., US Patent 6,716,274; 6,673,136.
4. Ko, F.K., Laurencin, C.T., Borden, M.D., and Reneker, D., Proceedings, Annual Meeting, Biomaterials
Research Society, San Diego, April 1998.
5. Rutledge, G., Fridrikh, S. et al., Texcomp 6, Philadelphia, PA, September 2002.
6. Ko, F.K., Nanofiber technology: bridging the gap between nano and macro world, in NATO ASI on
Nanoengineeered Nanofibrous Materials, Guceri, S., Gogotsi, Y. and Kuznetsov, V., NATO Series II,
Vol. 169, 2004.
7. Reneker, D.H. and Fong, H., Eds., Polymer nanofibers, Polymer Preprints, 44, September 2003,
American Chemical Society.
8. Li, W.J., Laurencin, C.T., Caterson, E.J., Tuan, R.S., and Ko, F.K., Electrospun nanofibrous structure:
a novel scaffold for bioengineering, J. Biomed. Mater. Res., 58, 613–621, 2002.
9. Boland, E.D., Simpson, D.G., Wnek, G. E., and Bowlin, G.L., Electrospinning of biopolymers for tis-
sue engineering scaffolds, Polymer Pre-preprints, 44, 92, 2003.
10. Kwoun, J., Lec, R.M., Han, B., and F. Ko. K., A novel polymer nanofiber interface for chemical sen-
sor applications, Proceedings of the IEEE International Frequency Control Symposium, May 7–9,
2000, Kansas City, MS, pp. 52–57.
11. MacDiarmid, A.G., Jones, W.E., Jr., Norris, I.D., Gao, J., Johnson, A.T., Jr., Pinto, N.J., Hone, J., Han,
B., Ko, F.K., Okusaki, H., and Llanguno, M., Electrostatically-generated nanofibers of electronic poly-
mers, Synth. Met., 119, 27–30, 2001.
12. Lifshutz, N. and Binzer, J.C., Nanofiber coated filter materials for automotive and other applications,
Proceedings of First International Congress on Nanofiber Science and Technology, June 28, 2004, The
Society of Fiber Science and Technology, Tokyo, Japan, pp. 54–62.
13. Ko, F. K., Yang, H., Argawal, R., and Katz, H., Electrospinning of improved CB protective fibrous
materials, Proceedings, March 30–31, 2004, Techtextil, Atlanta.
14. Vacanti, C.A. and Mikos, A.G., Letter from the editors, Tiss. Eng., 1, 1, 1995.
15. Gibson, P.W., Schreuder-Gibson, H.L., and Rivin, D., Electrospun fiber mats:transport properties,
AIChE J., 45, 190–195, 1999.
16. Ijimi, S., Helical microtubules of graphitic carbon, Nature, 354, 53–58, 1991.
17. Treacy, M.M., Ebbesen, T.W., and Gibson, J.M., Nature, 381,67–80, 1996.
18. Wang, E.W., Sheehan, P.E., and Lieber, C.M., Nanobeam mechanics: elasticity, strength and tough-
ness of nanorods and nanotubes, Science, 277, 1971–1975, 1997.
19. Lu, J.P., Elastic properties of single and multilayered nanotubes. Phys. Rev. Lett., 179, 1297, 1997.
20. Peebles, L.H., Carbon Fibers: Formation, Structure, and Properties. CRC Press, Boca Raton, FL, 1995.
21. Ko, F., Gogotsi, Y., Ali, A., Naguib, N., Ye, H., Yang, G., Li, C., and Willis, P., Electrospinning of con-
tinuous carbon nanotube-filled nanofiber yarns, Adv. Mater., 15, 1161–1165, 2003.
22. Ko, F.K., Lam, H., Titchenal, N., Ye. H., and Gogotsi, Y., Co-electrospinning of carbon nanotube rein-
forced nanocomposite fibrils, in Science and Technology of Electrospinning, Reneker, D. and Fong,
H., Eds., Wiley, New York, chapter X, in press.

Copyright 2006 by Taylor & Francis Group, LLC

You might also like