You are on page 1of 20

REVIEW

Electrospinning of Nanofibers:
Reinventing the Wheel?**
By Dan Li and Younan Xia*
Electrospinning provides a simple and versatile method for generating
ultrathin fibers from a rich variety of materials that include polymers,
composites, and ceramics. This article presents an overview of this
technique, with focus on progress achieved in the last three years. After a
brief description of the setups for electrospinning, we choose to concentrate on the mechanisms
and theoretical models that have been developed for electrospinning, as well as the ability to
control the diameter, morphology, composition, secondary structure, and spatial alignment of
electrospun nanofibers. In addition, we highlight some potential applications associated with the
remarkable features of electrospun nanofibers. Our discussion is concluded with some personal
perspectives on the future directions in which this wonderful technique could be pursued.

1. Introduction
One-dimensional (1D) nanostructures have been a subject
of intensive research due to their unique properties and
intriguing applications in many areas.[1] A large number of
synthetic and fabrication methods have already been demonstrated for generating 1D nanostructures in the form of fibers,
wires, rods, belts, tubes, spirals, and rings from various materials.[2] Among these methods, electrospinning (a drawing process based on electrostatic interactions) seems to provide the
simplest approach to nanofibers with both solid and hollow
interiors that are exceptionally long in length, uniform in
diameter, and diversified in composition.[3] Unlike other
methods for generating 1D nanostructures, the formation of a
thin fiber via electrospinning is based on the uniaxial stretching (or elongation) of a viscoelastic jet derived from a polymer solution or melt. This technique is similar to the commercial processes for drawing microscale fibers except for the use

[*] Prof. Y. Xia, Dr. D. Li


Department of Chemistry, University of Washington
Seattle, WA 98195-1700 (USA)
E-mail: xia@chem.washington.edu
[**] This work has been supported in part by AFOSR through the MURI
(awarded to the University of Washington) and DURINT (awarded
to SUNY Buffalo) programs, a Career Award from the NSF (DMR9983893), and a research fellowship from the David and Lucile
Packard Foundation. Y.X. is a Camille Dreyfus Teacher Scholar
(2002) and an Alfred P. Sloan Research Fellow (2000). We thank
Jesse McCann for his critical reading and correction of this manuscript.

Adv. Mater. 2004, 16, No. 14, July 19

of electrostatic repulsions between surface charges (rather


than a mechanical or shear force) to continuously reduce the
diameter of a viscoelastic jet or a glassy filament.[4] Compared
with mechanical drawing, electrostatic spinning is better suited for generating fibers with much thinner diameters, since
the elongation can be accomplished via a contactless scheme
through the application of an external electric field. Like
mechanical drawing, electrospinning is also a continuous
process and therefore should work well for high-volume production.
The electrospinning technique may be considered as a variant of the electrostatic spraying (or electrospray) process.[5]
Both of these techniques involve the use of a high voltage to
induce the formation of a liquid jet. In electrospray, small
droplets or particles are formed as a result of the varicose
break-up of the electrified jet that is often present with a
solution of low viscosity. In electrospinning, a solid fiber is
generated as the electrified jet (composed of a highly viscous
polymer solution) is continuously stretched due to the electrostatic repulsions between the surface charges and the evaporation of solvent. Electrospray has found widespread use in
many areas including mass spectrometry, painting, inkjet
printing, and manufacturing of particles with various sizes and
compositions.[5] The first patent that described the operation
of electrospinning appeared in 1934, when Formalas disclosed
an apparatus for producing polymer filaments by taking
advantage of the electrostatic repulsions between surface
charges.[6] Up until 1993, this technique had been known as
electrostatic spinning, and there were only a few publications
dealing with its use in the fabrication of thin fibers.[7] In the

DOI: 10.1002/adma.200400719

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1151

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

early 1990s, several research groups (in particular, the Reneker group at the University of Akron) revived interest in this
technique by demonstrating the fabrication of thin fibers from
a broad range of organic polymers. At this time the term
electrospinning was coined and is now widely used in the
literature.[3,8] These timely demonstrations triggered a lot of
experimental and theoretical studies related to electrospinning. It is notable that the number of publications in this field
(see Fig. 1) has been increasing exponentially in the past few
years, on account of the remarkable simplicity, versatility, and
potential uses of this technique.

Figure 1. The annual number of publications on the subject of electrospinning, as provided by the search engine of SciFinder Scholar. For
2004, there are already 71 publications before April 30.

The major objectives of this article are the following: a) to


briefly discuss the mechanism behind electrospinning and a
number of setups for generating ultrathin fibers with controllable diameters and compositions; b) to fully demonstrate the
potential of this technique in producing nanofibers with wellcontrolled structures, morphologies, and functionalities; c) to
address experimental issues related to the organization of
electrospun fibers into hierarchically structured assemblies;
and d) to assess a range of intriguing applications associated
with both individual electrospun nanofibers and their assemblies. Because the unique features and applications of electrospinning have recently been reviewed in a number of publications,[3] here we only intend to provide an overview of the
most recent activities in this area, with focus on the progress
in the past three years. The main text of this article is organized into four sections: the first section gives a brief introduction to the principle and various setups for electrospinning;
the second section provides a comprehensive account of the
strategies demonstrated for controlling the diameter, composition, structure, morphology, and orientation of electrospun
fibers; the third section discusses recent developments related
to the alignment and assembly of electrospun nanofibers; and
the fourth section summarizes some unique features/properties that distinguish electrospun nanofibers from 1D nanostructures prepared using other methods, as well as some
typical and potential applications of electrospun nanofibers.
Finally, a personal perspective on the future of this technique
is given.

Younan Xia was born in Jiangsu, China, in 1965. He received a B.S. degree in chemical physics from
the University of Science and Technology of China (USTC) in 1987, and then worked as a graduate
student studying nonlinear optical materials for four years at the Fujian Institute of Research on the
Structure of Matter, Academia Sinica. He came to the United States in 1991, received an M.S. degree
in inorganic chemistry from the University of Pennsylvania (with Professor Alan G. MacDiarmid)
in 1993, and a Ph.D. degree in physical chemistry from Harvard University (with Professor George
M. Whitesides) in 1996, after which he stayed at Harvard and worked as a postdoctoral fellow with
Profs. George M. Whitesides and Mara Prentiss. He moved to Seattle in 1997 and started as an
Assistant Professor of Chemistry at the University of Washington. He was promoted to the rank of
tenured Associate Professor in 2001 and to the rank of Professor in 2003. His research interests
include nanostructured materials, self-assembly, photonic crystals, colloidal chemistry, microfabrication, surface modification, electrospinning, conducting polymers, microfluidic and
microanalytical systems, and novel devices for photonics, optoelectronics, and displays.

Dan Li was born in Sichuan, China, in 1971. He received a Ph.D. degree in Materials Physics and
Chemistry from the University of Electronic Science and Technology of China (with Prof. Yanrong
Li and Prof. Yadong Jiang) in 1999. Subsequently, he spent two years as a postdoctoral fellow with
Prof. Xin Wang and one year as Associate Professor at Nanjing University of Science and
Technology, China. He joined Prof. Younan Xia's group at the University of Washington in Seattle
as a postdoctoral fellow in 2002. His research interests include electrospinning nanostructured
materials, conjugated organic polymers, and nanostructure-based electronic devices.
1152

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

2. Electrospinning: Setups and Mechanisms


2.1. The Basic Setup for Electrospinning
Figure 2 shows a schematic illustration of the basic setup
for electrospinning. It consists of three major components: a
high-voltage power supply, a spinneret (a metallic needle),

woven mat, see the inset of Figure 2 for the scanning electron
microscope (SEM) image of a typical sample. With the use of
this relatively simple and straightforward technique, more
than 50 different types of organic polymers have already been
processed as fibers with diameters ranging from tens of nanometers to a few micrometers.[3]

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

2.2. How Electrospinning Works


Although the setup for electrospinning is extremely simple,
the spinning mechanism is rather complicated. As in electrospray, electrospinning also involves complex electrofluid
mechanical issues. Before 1999, the formation of ultrathin
fibers by electrospinning was often ascribed to the splitting or
splaying of the electrified jet as a result of repulsion between
surface charges.[3a] Recent experimental observations demonstrate that the thinning of a jet during electrospinning is
mainly caused by the bending instability associated with the
electrified jet.[4,10] Figure 3A shows the photograph of a spinning jet.[10b] It is obvious that the jet was initially a straight
line and then became unstable. It appears that the cone-

Figure 2. Schematic illustration of the basic setup for electrospinning.


The insets show a drawing of the electrified Taylor cone and a typical
SEM image of the nonwoven mat of poly(vinyl pyrrolidone) (PVP) nanofibers deposited on the collector.

and a collector (a grounded conductor). Direct current (DC)


power supplies are usually used for electrospinning although
the use of alternating current (AC) potentials is also feasible.[3,9] The spinneret is connected to a syringe in which the
polymer solution (or melt) is hosted. With the use of a syringe
pump, the solution can be fed through the spinneret at a constant and controllable rate. When a high voltage (usually in
the range of 1 to 30 kV) is applied, the pendent drop of polymer solution at the nozzle of the spinneret will become highly
electrified and the induced charges are evenly distributed
over the surface. As a result, the drop will experience two
major types of electrostatic forces: the electrostatic repulsion
between the surface charges; and the Coulombic force exerted
by the external electric field. Under the action of these electrostatic interactions, the liquid drop will be distorted into a
conical object commonly known as the Taylor cone (see the
inset of Fig. 2).[3,4] Once the strength of electric field has surpassed a threshold value, the electrostatic forces can overcome the surface tension of the polymer solution and thus
force the ejection of a liquid jet from the nozzle. This electrified jet then undergoes a stretching and whipping process,
leading to the formation of a long and thin thread. As the
liquid jet is continuously elongated and the solvent is evaporated, its diameter can be greatly reduced from hundreds of
micrometers to as small as tens of nanometers. Attracted by
the grounded collector placed under the spinneret, the
charged fiber is often deposited as a randomly oriented, nonAdv. Mater. 2004, 16, No. 14, July 19

Figure 3. Photographs illustrating the instability region of a liquid jet


electrospun from an aqueous solution of poly(ethylene oxide) (PEO).
The capture time was on two different scales: A) 1/250 s, and B) 18 ns,
respectively. Note that the path of the jet shown in B has been traced to
improve the visibility. These two figures were adapted from [10b] with
permission. Copyright Elsevier Science, 2001.

shaped, instability region is composed of multiple jets. However, a closer examination using high-speed photography
(Fig. 3B) establishes that the conical envelope contains only a
single, rapidly bending or whipping thread. In some cases,
splaying of the electrified jet might also be observed, though
it was never a dominant process during spinning.[10] The frequency of whipping is so high that conventional photography
cannot properly resolve it, giving the impression that the original liquid jet splits into multiple branches as it moves toward
the collector.
Based on experimental observations and electrohydrodynamic theories, mathematical models have been developed by
several groups to investigate the electrospinning process.

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1153

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

Reneker and co-workers treated the charged liquid jet as a


system of connected, viscoelastic dumbbells and provided a
good interpretation for the formation of bending instability.[4a,10a] They also calculated the three-dimensional trajectory
for the jet using a linear Maxwell equation and the computed
results were in agreement with the experimental data. Rutledge and co-workers considered the jet as a long, slender object and thereby developed a different model to account for
the electrospinning phenomenon.[10b,11] Their experimental
and theoretical studies clearly showed that the spinning process only involves whipping (rather than splaying) of a liquid
jet. The whipping instability is mainly caused by the electrostatic interactions between the external electric field and the
surface charges on the jet. The formation of fibers with fine
diameters is mainly achieved by the stretching and acceleration of the fluid filament in the instability region. The same
group further showed that the model could be extended to
predict the saturation of whipping amplitude, as well as the
diameter of resultant fibers.[11c] In some related studies, Feng
proposed another model to describe the motion of a highly
charged liquid jet in an electric field, and the role of nonlinear
rheology in the stretching of an electrified jet was also examined.[12] All these studies provide a better understanding of
the mechanism responsible for electrostatic spinning process.
More importantly, they may assist experimentalists enormously in the design of new setups that may provide a better
control over the diameter and structure of electrospun nanofibers.

2.3. Modification of the Setup for Electrospinning


The basic setup for electrospinning is so simple that it has
already found widespread use in many research laboratories.
In order to further control the electrospinning process and
thus to tailor the structures of resultant fibers, the setup (in
particular, the collector and spinneret) has also been modified
in a number of ways. For example, a rotating drum has been
introduced by Reneker and co-workers to collect electrospun
fibers as relatively uniform mats.[13] It has also been demonstrated by several groups that the orientations of electrospun
fibers could be readily controlled by modifying the design or
layout of the collector electrodes. A more detailed discussion
with regard to the spatial alignment of electrospun fibers can
be found in Section 4.
In general, electrospinning is a process with low productivity because the polymer solution has to be fed at relatively
slow rates (usually less than 1 mL h1) in order to obtain ultrathin fibers. To solve this problem, an array of multiple needles
has recently been demonstrated by Chu and co-workers as the
spinneret for use with electrospinning.[14] In this case, electric
field interference between adjacent nozzles may arise, and the
configuration for the arrayed needles needs to be carefully
designed to ensure that the electric field strength at the tip of
each needle is identical in order to produce samples contain-

1154

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ing uniform fibers. In the same vein, a number of other methods can also be explored to fabricate the arrayed capillaries.
For example, one can take a commercial alumina membrane,
sputter one of its surfaces with gold and glue the other side to
the tube of a syringe to form an array of nozzles whose diameters can be readily varied in the range from hundreds of nanometers to tens of micrometers. For this system, it is necessary
to derivatize the gold surface of the membrane with an alkanethiolate monolayer terminated in CH3 or CF3 group to
confine the liquid within each individual pore and thus to
eliminate the coalescence of individual droplets.[15] In another
approach, silicon micromachining can be used to fabricate
porous membranes with patterned arrays of micrometer-sized
channels.[16] By varying the mask design, one can also easily
control the spacing between adjacent holes to avoid coalescence of liquid droplets coming out from different pores. In a
third approach, the rigid needles made of stainless steel that
are commonly used for electrospinning can be replaced with
flexible, polyimide-coated glass capillaries commercially
available for use in capillary electrophoresis.[17] The diameters
of such glass capillaries can be varied from a few to several
hundred micrometers. They can also be easily assembled into
a hexagonal array that may contain tens to hundreds of capillaries. By connecting each capillary to a different solution, it
will be possible to simultaneously electrospin nanofibers with
a range of different compositions and to obtain assemblies of
nanofibers that will exhibit various unique combinations of
functionalities. It is believed that the adoption of an array of
nozzles can significantly increase the production rate while
simultaneously increasing considerably the compositional and
structural complexities associated with electrospun nanofibers
and their assemblies.
In addition to the use of capillaries having hollow interiors,
spinnerets can also be fabricated with solid tips. For example,
Craighead and co-workers have demonstrated that a microfabricated, pyramidal silicon tip could serve as the spinneret
for electrospinning when it was integrated with a microfluidic
device.[18] A combination of electrospinning and microfluidics
may provide an exciting opportunity to directly process functional materials on miniaturized chips. In a related study,
Kessick and Tepper demonstrated that the setup for electrospinning could be miniaturized to produce single polymer
fibers (or their networks) on the surface of a microchip without using high voltage, a syringe pump, or a metal needle.[19]
In this case, they placed a drop of polymer solution on the surface of an interdigitated array of metal electrodes and found
that fibers could be formed between drops sitting on adjacent
electrodes if a potential difference was applied between the
two sets of electrodes or if the drops were charged in advance
via electrospray. These two demonstrations imply that electrospinning can probably be integrated with the conventional
lithographic techniques to form a new fabrication platform
for generating patterned microstructures from various materials and on a broad range of length scales.
Similar to electrospray, the conventional setup for electrospinning also involves the use of a single capillary as the spin-

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

neret, and thus is only suitable for generating fibers with one
particular composition in each run of fabrication. To overcome this limitation, a number of new designs for the spinneret have recently been demonstrated. In one approach, Wilkes
and Balkus Jr. and their co-workers showed that a spinneret
containing two needles (attached to each other in a side-byside fashion) could be used to electrospin composite nanofibers.[20] In another approach, a spinneret consisting of two
coaxial capillaries has been demonstrated by several groups
to fabricate fibers with core/sheath or hollow structures.[21] In
particular, the spinneret design demonstrated by Xia and Li is
simple enough that it can be conveniently fabricated at essentially no cost by any research group (see Sec. 3.3.2 for a
detailed discussion). Larsen et al. have also demonstrated that
the use of a coaxial spinneret could provide a good control
over the structure and morphology of resultant fibers, and
prevent blockade of the nozzle by flowing a suitable solvent
vapor through the outer capillary during an electrospinning
process.[22]

3. How to Control Electrospun Nanofibers


3.1. Control of Morphology and Diameter

of at least three forces. As driven by minimization of surface


area, surface tension tends to convert the liquid jet into one
or many spherical droplets (Rayleigh instability[24]). On the
other hand, the electrostatic repulsion between charges on the
jet surface tends to increase the surface area, and thus favors
the formation of a thin jet rather than beads. Viscoelastic
force also resists rapid changes in shape and supports the
formation of fibers with smooth surfaces. In general, the formation of beads can be eliminated whenever the influence of
surface tension is suppressed by the effects of the last two
forces. As demonstrated by these authors, the density of beads
in PEO fibers decreased or disappeared if more viscous solutions (i.e., higher concentrations of PEO) were used. Addition
of salts to increase the net charge density or use of solvents
with lower surface tensions could also lead to the effective
elimination of beads. Xia and Li have also studied the dependence of morphology of poly(vinyl pyrrolidone) (PVP) nanofibers on the concentration of the electrospinning solution.
Figure 4 shows several representative SEM images of PVP
nanofibers that were electrospun from solutions of various
concentrations. It can be clearly seen that the density of beads
was dramatically reduced as the concentration of PVP was
increased. When polymer solutions of the same concentration
were used, the introduction of a small amount of salt such as
tetramethylammonium chloride could yield bead-free nanofibers with diameters as small as 90 nm. Besides concentration, Tan and co-workers have also found that the formation

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

The morphology and diameter of electrospun fibers are dependent on a number of processing parameters that include:
a) the intrinsic properties of the solution such as the type of polymer, the
conformation of polymer chain, viscosity (or concentration), elasticity, electrical conductivity, and the polarity and
surface tension of the solvent; and
b) the operational conditions such as
the strength of applied electric field, the
distance between spinneret and collector, and the feeding rate for the polymer
solution.[3] In addition to these variables, the humidity and temperature of
the surroundings may also play an important role in determining the morphology and diameter of electrospun
nanofibers (see Sec. 3.3.3).
The presence of beads in electrospun
fibers is a common problem.[23] With
poly(ethylene oxide) (PEO) as a model
system, Reneker and co-workers have
systematically investigated the influence
of solution properties on the density of
beads contained in the electrospun
fibers.[23a] It was found that the viscosity,
surface tension, and the density of net
charges carried by the liquid jet all
Figure 4. SEM images of poly(vinyl pyrrolidone) (PVP) nanofibers that were electrospun from PVP
influenced the morphology of resultant
solutions in a mixture of ethanol and water (16:3 by volume). The weight percentage of PVP in the
fibers. However, formation of beads
solution was: A) 3, B) 5, C) 7, and D) 5. In preparation of sample D, 0.35 mg mL1 of tetramethylammonium chloride was added to the solution before electrospinning.
could be attributed mainly to the action

Adv. Mater. 2004, 16, No. 14, July 19

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1155

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

of beads in the fibers had some correlation to the spinning


voltage.[23c] In general, increasing voltage to a certain level
would change the shape of the pendent drop from which the
jet originated so that a stable, conical shape could not be
achieved. As a result, the stability of the liquid jet would be
weakened, which might lead to an increase in the density of
beads in the electrospun fibers.
Fibers obtained by electrospinning are usually circular in
cross-section.[8] Other shapes, in particular ribbon-like structures with rectangular cross-sections, have also been observed.[25] It is believed that the formation of a skin on the
surface of liquid jet (due to rapid evaporation of solvent) and
the subsequent collapse of the skin might be responsible for
the emergence of a ribbon-like morphology. For electrospun
fibers with circular cross-sections, the control of diameters has
been discussed in a number of publications.[26] It has been
established that the major factors that affect the diameter of
electrospun fibers include the concentration of polymer, the
electrical conductivity of the solution, the electric field
strength, and the feeding rate for the solution. In general, the
fibers become thicker as more concentrated solutions are
used. If the conductivity of a solution is increased by adding
some salts, the diameters of the resultant fibers can be significantly reduced. A higher feeding rate for the solution always
leads to the formation of thicker fibers. The influence of spinning voltage seems to be poorly established. Some studies
showed that the diameter decreased with increasing the spinning voltage, while others showed opposite trends. A better
parameter that should be used here might be the strength of
electric field, rather than the voltage. On the theoretical side,
Rutledge and co-workers have recently proposed a model to
account for the influence of various processing parameters on
the size of electrospun fibers.[11c] In this model, the final diameter of a spinning jet strongly depends on the interplay
between surface tension and electrostatic repulsions. The balance point might be related to the flow rate, the strength of
electric field, and the surface tension of liquid phase. The
diameters predicted using this model were in agreement with
experimental results.

3.2. Control of Chemical Composition


Early work on electrospinning mainly dealt with conventional organic polymers that could be synthesized with sufficiently high molecular weights and could be dissolved in
appropriate solvents. In an effort to greatly expand the functionality and thus the scope of applications associated with
fibrous structures, a variety of methods have recently been
developed to increase the diversity of materials that can be
adapted for use with electrospinning. This section only discusses four of these methods that were capable of generating
nanofibers with a range of chemical compositions, and therefore various electronic, magnetic, optical, and biological properties. It should also be possible to incorporate a combination

1156

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

of these functionalities into the same sample to obtain multifunctional nanofibers.


3.2.1. Functional Polymers and Their Blends
A number of functional polymers with applications have
been directly electrospun into nanofibers. For example,
MacDiarmid et al. showed that polyaniline, a conductive polymer, could be prepared as fibrous structures by electrospinning when sulfuric acid was used as the solvent.[27] Poly(vinylidene fluoride) (PVDF), a well-known piezoelectric polymer,
has been electrospun as nanofibers by several research
groups.[25a,28] Samuelson and co-workers have successfully
processed a fluorescent polymer into nanofibers by electrospinning and then demonstrated their application in optical
sensing.[29] In addition, Balkus Jr. and co-workers have electrospun poly[2-methoxy-5-(2-ethylhexyloxy)-1,4-phenylenevinylene] (or MEH-PPV), one of the best-established photoor electroluminescent polymers, into thin fibers.[20b] In this
case, a large amount of leaf-like structures also existed in the
resultant fibers. Most recently, Smith and co-workers have
demonstrated that Pt(NH2dmoc)4(PtCl4) (a metal-containing,
semiconducting polymer) could also be electrospun as long
nanofibers, which were then exploited as active components
to fabricate field-effect transistors (FETs).[30]
Intrigued by their potential applications as scaffolds in cell
biology and tissue engineering, a large number of biodegradable polymers that include poly(caprolactone), poly(L-lactide)
(PLA), and poly(glycolide) have been directly electrospun
into nanofibers by Kaplan and several other groups.[26,31] In
addition to these synthetic organic polymers, natural biopolymers such as DNA, silk fibroin, human or bovine fibrinogens,
dextran, collagens, and even viruses, have also been successfully used for electrospinning.[32] In general, these biomacromolecules observe the same rules that have been derived from
studies on conventional synthetic polymers.
As limited by their molecular weights and/or solubilities,
there are many functional polymers that are not suitable for
use with electrospinning. One of the most effective strategies
for solving this problem is to blend them with polymers that
are well-suited for electrospinning. Based on this approach,
Kaplan and co-workers have successfully fabricated proteincarrying fibers by adding the proteins to the solution of a
conventional polymer.[33] Nanofibers consisting of blends
between polyaniline (or polythiophene) with conventional
organic polymers have also been investigated by MacDiarmid
and other groups.[34] Blending was found to be fruitful in
improving some properties or applications associated with
nanofibers. For instance, Kim et al. demonstrated that the
physical and biological properties (e.g., biodegradation rate
and hydrophilicity) of PLA nanofibers could be finely tuned
by simply controlling the compositions of polymer blend
solutions used for electrospinning.[35] It is worth noting
that the polymers contained in an electrospinning solution
might separate into different phases as the solvent is evaporated.

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

3.2.2. Inorganic/Polymer Composites and Ceramics


Historically, electrospinning has been largely limited to the
fabrication of nanofibers from organic polymers due to the
stringent requirements on the viscoelastic behavior of a solution. Recent efforts by several groups established that conventional solgel precursor solutions could also be employed for
electrospinning to provide an effective route to composite
and ceramic nanofibers. For example, Larsen et al. have demonstrated the direct electrospinning of viscous inorganic sols
and successfully fabricated fibers consisting of TiO2/SiO2 and
Al2O3.[36] Using a similar approach, Santigano-Aviles and coworkers have also fabricated PbZrxTi1xO3 (PZT) fibers.[37]
Choi, Ramsier, and their co-workers have obtained SiO2
nanofibers.[38] In addition, Balkus Jr. and co-workers have
demonstrated the preparation of nanofibers from a series of
mesoporous molecular sieves by electrospinning solution containing appropriate sols and structure-directing agents.[39] The
key strategy to the success of these demonstrations was to
control the hydrolysis rate of a solgel precursor by adjusting
the pH value or aging conditions. It is important to form a
solution with viscoelastic behavior similar to that of a conventional polymer solution.
The fibers prepared via direct electrospinning of inorganic
sols were usually several hundred nanometers in diameter. It
might be possible to further reduce the size of these fibers but
such efforts will be hindered by the difficulty of precisely controlling the viscoelastic behavior of a sol-based gel system.
Xia and Li recently demonstrated a new approach in which
the solgel reaction took place mainly in the spinning jet
rather than in the stock solution.[40] In a typical procedure, a
solgel precursor such as a metal alkoxide was mixed with a
PVP solution in alcohol. The role of PVP in the solution was
to increase the viscosity and thus to control its viscoelastic
behavior. After the solution had been electrospun into a thin
jet, the metal alkoxide immediately started to hydrolyze by
reacting with the moisture in surrounding air to generate a
continuous gel network within the polymer matrix. As a
result, nanofibers consisting of an inorganic/polymer composite would be obtained as the immediate product. Ceramic
nanofibers could be readily obtained by removing the organic
phase via calcination of these composite fibers at elevated
temperatures in air. The diameter of these ceramic fibers
could be controlled from tens of nanometers to several hundred nanometers with relatively narrow size distribution. Figure 5 shows the SEM and transmission electron microscope
(TEM) images of two typical samples of anatase nanofibers
that were obtained by calcining electrospun nanofibers made
of PVP and amorphous titania. With the use of this simple
procedure, anatase nanofibers as thin as 20 nm could be readily fabricated (Fig. 5B). This method has also been successfully extended to process many other oxide ceramics into
nanofibers that include, for example, SiO2, SnO2, indium tin
oxide (ITO), GeO2, NiFe2O4, LiCoO2, and BaTiO3. Both
SEM and TEM analyses indicated that the ceramic nanofibers

Adv. Mater. 2004, 16, No. 14, July 19

Figure 5. SEM (A) and TEM (B) images of anatase nanofibers prepared
by electrospinning from ethanol solutions that contained 0.03 g mL1
poly(vinyl pyrrolidone) (PVP) and different amounts of titanium tetraisopropoxide: A) 1.5 and B) 0.25 g mL1, respectively. Both samples were
calcined in air at 500 C before imaging [40a].

were polycrystalline in structure, with domain size on the


scale of a few nanometers. It was also found that the nanocrystallites in these fibers were more uniform in dimension
than those prepared using a conventional solgel process,
because the moisture could quickly diffuse into the thin fibers,
causing a rapid and uniform hydrolysis for the alkoxide precursor.
In addition to PVP, poly(vinyl alcohol) (PVA), poly(vinyl
acetate), and PEO have also been used as the polymer matrices to host inorganic precursors by several groups.[41] A
number of oxides that include Al2O3, CuO, NiO, TiO2SiO2,
V2O5, ZnO, Co3O4, Nb2O5, MoO3, and MgTiO3 have been
fabricated as fibrous structures.[41] More interestingly, non-oxide ceramic nanofibers could also be prepared using this approach: for instance, Larsen et al. have fabricated SiC fibers
by electrospinning a sol solution containing Novolac resin and
tetraethylorthosilicate, followed by conversion of these fibers
to SiC using high-temperature pyrolysis.[36] All these studies
clearly demonstrate that electrospinning may provide a
powerful route to inorganic nanofibers that will find use in
many applications, for example, as components for structural
reinforcement, as active units in sensing, as membranes for
purification and separation, as supports in catalysis, and as
electrode materials in energy conversion or storage devices
that include batteries, dye-sensitized solar cells, and actuators.

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1157

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

3.2.3. Encapsulation of Functional Materials


A variety of functional components (e.g., nanoparticles,
nanowires, or molecular species) can be directly added to the
solution for electrospinning to obtain nanofibers with a diversified range of compositions and well-defined functionalities.
To this end, incorporation of nanoparticles such as titania, carbon black, silver, and iron oxides have all been demonstrated
(Fig. 6A).[42] Introduction of nanoparticles into polymer nanofibers have also been accomplished by adding appropriate
precursors to the solution for electrospinning, followed by

Figure 6. A) TEM image of poly(vinyl pyrrolidone) (PVP) nanofibers


whose interiors had been encapsulated with iron oxide nanoparticles by
directly adding this material to the electrospinning solution. B) High-resolution TEM image showing the alignment of CNTs in a poly(acrylonitrile) fiber with the diameter of about 50 nm. The inset shows a schematic illustration of the orientation of CNTs in a fiber. This figure was
adapted from [44a].

post-spinning treatment.[43] For instance, Greiner and coworkers added palladium diacetate to a PLA solution to
increase the electrical conductivity of this solution in an effort
to produce thinner nanofibers.[43a] Annealing of the as-spun
nanofibers at an elevated temperature led to the formation of
Pd nanoparticles within each fiber. Reneker and Hou have
also demonstrated the incorporation of iron salts into a poly(acrylonitrile) (PAN) solution to be used for electrospinning.[43b] Calcination of the resultant fibers in an appropriate
atmosphere resulted in the carbonization of PAN polymer
and the reduction of iron salts to iron nanoparticles. These
iron nanoparticles could be further used as catalysts to direct
the growth of carbon nanotubes (CNTs) on the carbon nanofibers. In general, it is non-trivial to achieve a homogeneous
1158

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

distribution for the encapsulated nanoparticles within each individual nanofiber.


Incorporation of relatively short nanowires or nanotubes
into nanofibers also represents an exciting direction of
research for electrospinning. For example, incorporation of
CNTs into electrospun nanofibers may greatly improve the
mechanical strength and electronic conductivity of the fibers.
This work has been carried out by a number of research
groups.[44] For example, Ko and co-workers have found that
the orientation of CNTs in polymer fibers was dependent on
the type of polymer matrix and how well the CNTs were dispersed in the polymer solution.[44a] Because single-walled
CNTs could be dispersed well in a N,N-dimethyl formamide
(DMF) solution of PAN, the CNTs could maintain their
straight morphology and be oriented parallel to the axial
direction of the electrospun PAN fibers (Fig. 6B). In contrast,
when CNTs were added to a solution of poly(lactic acid)
(PLA), the CNTs were distributed inhomogeneously within
the fibers in the form of highly tangled agglomerates. Cohen
and co-workers proposed a theoretical model to investigate
the behavior of rod-like particles in the electrospinning jet
and pointed out that it was feasible to align rod-like particles
along the jet line.[44b] They obtained PEO nanofibers loaded
with multi-walled CNTs and found that only a small portion
of CNTs appeared to be well-oriented along the fiber axis,
and the product exhibited a certain degree of tortuousity.
Most recently, Seoul et al. electrospun PVDF into nanofibers
together with CNTs as the fillers.[44c] They also compared the
conductivities of electrospun fibers and spin-coated thin films
and found that the fibers exhibited a higher percolation
threshold for the insulator-to-conductor transition.
In additional to nanometer-sized objects, drugs, dyes,
enzymes, DNA, and other molecular species could also be
readily incorporated into electrospinning solutions to produce
functional nanofibers.[45] More interestingly, if these additives
were not soluble in the solutions, they could also be introduced in the form of emulsions. For example, Wnek and coworkers demonstrated a two-phase electrospinning procedure, by which aqueous reservoirs could be encapsulated in
the fibers of poly(ethylene-co-vinyl acetate).[46] A rich variety
of functional components could be introduced into fibers in
this fashion. It is worth noting that the addition of any material to the spinning solution may cause some changes to its
properties (e.g., conductivity, viscosity, elastic strength, and
viscoelasticity) and thus alter the structure and morphology of
resultant fibers.[47]
3.2.4. Modification of Electrospun Nanofibers
The nanofibers prepared by electrospinning could be modified in a number of different ways to improve their properties
and/or to increase the diversity of materials that could be processed as fibrous nanostructures. Surface coating seems to
represent the simplest approach in this regard. For example,
Samuelson and co-workers have shown that the surfaces of
electrospun nanofibers could be deposited with fluorescent

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

molecules, conducting polymers, or titania nanoparticles via


electrostatic interactions or liquid-phase attachment.[48] Pinto
et al. have coated PAN or polyaniline fibers with sheaths of
nickel or gold through electroless deposition.[49] In addition,
physical or chemical vapor depositions could be exploited to
coat or decorate as-spun fibers with polymers, metals, or
ceramics.[50] To this end, Xia and co-workers demonstrated
that the surface of titania nanofibers could be readily decorated with nanoparticles of noble metals by taking advantage
of the photocatalytic activity of anatase.[51] Figure 7 shows the
SEM images of two typical samples where the surfaces of

mechanical drawing.[52] As a result, the degradation behaviors


of these mats could be greatly improved. For ceramic nanofibers, their mechanical and transport properties could be
enhanced by annealing the samples at an elevated temperature to induce the formation of crystallites having the right
phase structures and/or domain sizes.
Chemical modification provides by far the most versatile
route to nanofibers characterized by new compositions and
functionalities. In the simplest procedure, electrospun nanofibers could be thermally transformed into other materials
without changing their morphology via calcination at elevated
temperatures. Using this approach, many materials that cannot be directly used for electrospinning have been prepared
as thin fibers with controllable diameters and compositions.
The preparation of ceramic nanofibers described in Section 3.2.2. provides some good examples. Another good
example involves the synthesis of carbon nanofibers by carbonization of fibers electrospun from different types of precursor polymers.[53] In other demonstrations, Liu and Hsieh
introduced double-bond units into cellulose fibers via alkaline
deacetylation, followed by reaction with methacrylate chloride.[54] Such methacrylated cellulose fibers could be further
copolymerized with other types of monomers. Chemical
modifiers containing double-bond units could also be added
to the solution before electrospinning. In this case, chemical
reactions (e.g., crosslinking or grafting) could be accomplished during electrospinning by light-initiated polymerization or by post-spin treatments.[55] Kenawy and Abdel-Fattah
modified poly(vinyl phenol) fibers using a lithium salt to
improve their antimicrobial activity.[56] These demonstrations
clearly establish that the electrospun nanofibers provide a
unique platform for fabricating 1D nanostructures characterized by a broad range of compositions and functionalities.

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

3.3. Control of Secondary Structures


Figure 7. SEM images of anatase nanofibers whose surfaces had been
decorated with gold (A) and silver (B) nanoparticles via photocatalytic
reduction of HAuCl4 and AgNO3 in aqueous solutions, respectively.

anatase nanofibers had been deposited with gold and silver


nanoparticles via a photocatalytic reduction process. By varying the concentration of metallic salts and/or the exposure
time to UV light, it was possible to control the size and density of deposited metal nanoparticles. Through these modifications, the chemical, thermal, mechanical, liquid wetting, and
absorbent properties of the fibrous nanostructures could be
significantly altered without changing the morphology of original fibers. These demonstrations also open the door to add
new functionalities to nanofiber-based membranes and materials.
The properties of electrospun nanofibers can also be modified using a number of physical processes. For example, Chu
and co-workers found that the orientation of poly(glycolideco-lactide) chains in electrospun, nonwoven mats could be significantly changed when the samples were annealed under
Adv. Mater. 2004, 16, No. 14, July 19

Nanofibers prepared by electrospinning usually exhibit a


solid interior and smooth surface. Recent demonstrations
imply that nanofibers with some specific secondary (e.g., core/
sheath, hollow, and porous) structures could also be prepared
if appropriate processing parameters or new designs of spinnerets were employed. Here we only discuss three types of
such structures, with an emphasis on the mechanisms and
potential extensions.
3.3.1. Nanofibers with Core/Sheath Structures
With the use of conventional setup for electrospinning, it is
possible to fabricate core/sheath nanofibers when a polymer
solution containing two polymers that will phase separate as
the solvent is evaporated. For example, Sung and co-workers
observed the formation of core/sheath structures when a solution containing poly(carbonate) (PC) and poly(butadiene)
was electrospun into thin fibers.[57] Yarin, Greiner, and coworkers have recently demonstrated that core/sheath nano-

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1159

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

fibers could be fabricated by co-electrospinning two different


polymer solutions through a spinneret comprising two coaxial
capillaries (see Fig. 8A for a generic design).[21a] It was proposed that mixing of the two solutions could be prevented by
the low diffusion coefficients of polymer chains because the
spinning process took place very quickly and there would not
be enough time for the chains to be mixed before solidification. However, no information was provided with regard to
the yields of nanofibers with core/sheath structures, as well as
the uniformity of resultant fibers because the optical and
TEM images presented in their publication contained only
very short segments of individual nanofibers.
In a related work, Xia and Li systematically investigated
this new spinneret system,[21b] and found that the formation of
a compound liquid jet (and thus a core/sheath nanofiber) was
strongly affected by all processing parameters. For example,
the yield of this approach was dependent on the viscosities of
both core and sheath liquids. When the viscosity of liquid
ejected through the inner capillary was reduced to a certain
level, it would be impossible to stretch the core into a continuous, thin thread within the sheath. The presence of a solgel
ceramic precursor in the sheath solution was also necessary
for the formation of a stable, coaxial jet. In general, addition
of a solgel precursor such as Ti(OiPr)4 to the PVP solution
could greatly improve the stability of coaxial jets and thus
lead to the formation of core/sheath nanofibers with uniform
diameters. More importantly, the immiscibility of core and
sheath liquids played a crucial role in determining the formation of core/sheath jet with a continuous and uniform crosssection. It will be difficult or unfeasible to generate core/

sheath liquid jets with reasonable yields by electrospinning


two polymer solutions whose polymers or solvents are miscible. More work (including simulation) is still needed in order
to achieve a complete understanding and full control of this
new electrospinning system.
3.3.2. Nanofibers with Hollow Interiors

One-dimensional nanostructures with hollow interiors (e.g.,


nanotubes) are of fundamental and technical importance for a
range of applications such as nanofluidics and hydrogen storage. Many methods have been demonstrated for fabricating
such structures from a wealth of materials and on various
scales.[58] For example, nanofibers prepared by electrospinning have been used as sacrificial templates for generating
tubular fibers.[43a,50,59] In this method, additional coating and
etching steps are often required and the quality of the resultant tubes is strongly dependent on the yield and control of
each step. It is highly desirable to extend the capability of
electrospinning in an effort to directly generate hollow nanofibers in one step.
Recent efforts from at least two groups indicate that electrospinning could be directly utilized to fabricate hollow
nanofibers. Figure 8A shows a setup that was developed by
Xia and Li in their work on the electrospinning of nanofibers
with a core/sheath or hollow structure.[21b] In a typical procedure, two viscous but immiscible liquids, for example, mineral
oil and an ethanol solution containing PVP and Ti(OiPr)4)
were used as the source materials for core and sheath, and
they were ejected simultaneously through the inner and outer
capillaries to form a stable compound
jet. With the use of this co-spinning
process, core/sheath nanofibers made of
TiO2/PVP and mineral oil were
obtained as the immediately product.
Selective removal of the oil phase by
solvent extraction resulted in the formation of hollow fibers consisting of TiO2/
PVP composite walls (Fig. 8B). Polycrystalline, ceramic hollow nanofibers
could be obtained by removing both
PVP and oil through calcination of the
as-spun fibers at elevated temperatures
in air (see Fig. 8C,11D for typical TEM
and SEM images, respectively). By controlling the spinning conditions (e.g.,
electric field strength, concentration of
sheath liquid, and the feeding rates for
both liquids), the sheath thickness and
inner diameter of the nanofibers could
Figure 8. A) Schematic illustration of the setup used for direct fabrication of hollow nanofibers by
be varied in the range from tens of
electrospinning. It involves the use of a spinneret consisting of two coaxial capillaries, through
which mineral oil and an ethanol solution containing poly(vinyl pyrrolidone) (PVP) and titanium
nanometers to several hundred nanotetraisopropoxide were simultaneously ejected to form a compound jet. B) TEM image of the asmeters. The instability problem assospun hollow nanofibers whose walls were made of a composite of PVP and amorphous TiO2. The
ciated with polymer hollow structures
oil cores had been extracted with octane. C) TEM image of anatase hollow nanofibers that were
was also solved by introducing an inorobtained by calcining the composite nanotubes at 500 C in air. These figures were adapted from
[21b] with permission. Copyright American Chemical Society, 2004.
ganic solgel precursor into the spinning

1160

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

solution. The strength and stability of the resultant tubular


nanostructures could be significantly improved via the in situ
formation of a gel network in the polymer sheath during the
spinning process. Most recently, the same group also demonstrated that functional nanoparticles or molecular species
could be directly incorporated into the hollow interiors by
pre-dissolving these functional materials into the core liquid.
These demonstrations imply that electrospinning is a simple
and versatile technique capable of fabricating nanostructures
with complex functionalities and hierarchical structures.
With the use of a similar setup, Larsen and co-workers independently demonstrated that polymer-free, inorganic nanotubes (e.g., consisting of SiO2 and ZrO2) could be fabricated
by co-electrospinning an aged inorganic sol and an immiscible
(or poorly miscible) liquid such as olive oil or glycerin, followed by selective removal of the inner liquid.[21c] The resultant hollow fibers had a mean inner diameter on the scale of
500 nm and shell thickness around 70 nm. It was suggested
that these dimensional parameters could be further varied in
a controllable fashion by adjusting experimental conditions
such as electric field strength and liquid flow rates. Based on
these two demonstrations, it might also be possible to fabricate hollow nanofibers with multiple walls by using spinneret
composed of more than two coaxial capillaries.

3.3.3. Nanofibers with Porous Structures


The surface area of a nanofiber can be greatly increased
when its structure is switched from a solid to a porous one.
Increase of surface areas is beneficial to many applications
that include catalysis, filtration, absorption, fuel cells, solar
cells, batteries, and tissue engineering. Two slightly different
approaches have been reported for introducing a porous
structure into the bulk of an electrospun nanofiber. One of
them was based on the selective removal of a component from
nanofibers made of a composite or blend material. The other
one involved the use of phase separation of different polymers
during electrospinning under the application of proper spinning parameters. Wendorff and co-workers investigated the
structural changes for fibers consisting of a PLA/PVP blend
when one of the two components was selectively removed.[60]
It was found that the fibers became highly porous in structure
when equal amounts of the two polymers were loaded into
the electrospinning solution. If the amount of one polymer
was much higher than the other one, the fibers remained compact in structure after one of the polymer phases had been
removed. This morphological change was believed to result
from rapid phase separation and solidification in the spinning
jet. The same group also found that porous fibers of PLA and
PC could be directly obtained if the spinning parameters and
the solvent system were judiciously selected.[61] Figure 9A
shows an SEM image of porous PLA fibers that were
obtained by electrospinning its solution in dichloromethane.
It was possible to vary the pore size and density by controlling
the processing parameters.
Adv. Mater. 2004, 16, No. 14, July 19

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

Figure 9. A) SEM image of porous poly(L-lactide) (PLA) fibers fabricated


by electrospinning a solution of PLA in dichloromethane (adapted from
[60]). B) SEM images of porous anatase nanofibers fabricated by co-electrospinning an ethanol solution of Ti(OiPr)4/PVP and a poly(styrene)
(PS) solution in a mixture of N,N-dimethyl formamide (DMF) and tetrahydrofuran (THF) (1:1 by volume) through a coaxial, two-capillary
spinneret. The sample had been calcined in air to remove both PS and
PVP. The inset shows a cross-sectional SEM image of the nanofiber,
clearly indicating the formation of a mesoporous structure. This figure
was adapted from [21b] with permission. Copyright American Chemical
Society, 2004.

Rabolt and co-workers have also observed the formation of


pores in the electrospinning of other polymers such as PC,
poly(methyl methacrylate) (PMMA), and polystyrene (PS).
They found that the solvent vapor pressure and the humidity
in atmosphere strongly affected the formation of pores.[62] It
was suggested that the cooling effect due to rapid evaporation
of a highly volatile solvent might induce the polymers to separate into different phases in the liquid jet. Because of evaporative cooling and condensation, water droplets could also be
formed within the fibers to promote the formation of porous
nanofibers. A higher humidity level was found to cause an increase in the density and size for the pores. In addition to solutions of polymer blends, it should be possible to form porous
nanofibers by electrospinning solutions containing organic
block copolymers.
Most recently, Xia and Li demonstrated that porous ceramic nanofibers could be fabricated by electrospinning with the
use of a coaxial, double-capillary system.[21b] In this case, the
PS solution in a mixture of DMF and tetrahydrofuran (THF)
was used as the core liquid and a PVP/Ti(OiPr)4 solution in
ethanol was used as the sheath liquid. Although the two polymers (PS and PVP) are immiscible, the solvents are miscible.

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1161

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

It was found that these two liquids could be partially mixed in


the electrospinning process because the mutual diffusion of
the two solvents could bring PS and PVP/Ti(OiPr)4 together
to form a more or less homogeneous mixture in certain
regions of each liquid jet. As the solvents evaporated quickly
in the next steps, these two polymer phases were separated to
generate nanoscale domains of PS embedded in a continuous
TiO2/PVP matrix. Once the PS phase had been removed by
calcination (along with the PVP phase), the resultant fibers
became highly porous in the PS-rich regions. Figure 9B shows
the SEM image of such a sample. It can be seen that each
fiber is composed of segments with different contrasts under
SEM: the dark regions are solid anatase structures with
smooth surfaces (as revealed by high-magnification SEM
imaging) while the bright regions are highly porous (see the
inset for a magnified SEM image). This and other demonstrations suggest that it is feasible to control the secondary structure of electrospun fibers by varying the polymers contained
in the core and sheath solutions. When used as the building
blocks to construct a thin film or membrane, the mesoscale
pores in individual nanofibers will contribute greatly to the
surface area and will usually result in the formation of a hierarchically porous structure.

4. Alignment of Electrospun Nanofibers

to each other along the edge. These authors also simulated


the electrostatic field of this configuration and revealed that
the field strength increased dramatically near the edge of the
disk. Because of the strong electrostatic attraction, the
charged fibers were continuously wound on the edge when
the disk was rotating at a relatively high speed. They further
demonstrated that nanofiber crossbars could be readily fabricated using this collector.[64b] With the use of a similar setup,
Natarajan, Xu, and their co-workers have also fabricated
well-aligned nanofibers.[65] In addition to drums, metal or
wooden frames have been explored by Vaia, Wendorff, and
their co-workers to collect electrospun nanofibers as more or
less aligned arrays.[66] Electrospun fibers have also been
aligned into parallel arrays using a multiple-field technique.[67]
Although these last two methods were somewhat successful in
aligning the nanofibers, it is still not clear what mechanisms
were responsible for the observed alignment effects.

4.2. Collector Based on a Pair of Split Electrodes


Xia and Li recently demonstrated that the geometrical configuration of a conductive collector had a profound effect on
the orientation of electrospun nanofibers.[40b,68] By using a collector consisting of two conductive strips separated by a void
gap of variable widths (up to several centimeters, see Fig. 10A
for a generic configuration), electrospun fibers could be uni-

For many applications, it is necessary to control the spatial


orientation of 1D nanostructures. In the fabrication of electronic and photonic devices, for example, well-aligned and
highly ordered architectures are often required.[63] Even for
application as simple as fiber-based reinforcement, it is also
critical to control the alignment of fibers. Because of the
bending instability associated with a spinning jet, electrospun
fibers are often deposited on the surface of collector (a single
piece of conductive substrate) as randomly oriented, nonwoven mats. In the past several years, a number of approaches
(mainly relying on the modification of jet movement by controlling the distribution of electric field) have been demonstrated to directly collect electrospun nanofibers as uniaxially
aligned arrays. Here we only discuss two of these approaches
that involved the use of a rotating drum (or frame) or a pair
of split electrodes as the collector.

4.1. Collector Based on a Rotating Drum or Frame


Alignment of electrospun fibers was observed by several
groups when a cylinder with high rotating speed was used as
the collector.[13,18a] Air flow may also favor the orientation of
fibers along the winding direction.[53a] However, the degree of
orientation for nanofibers collected by rotating drums was far
from perfect. Zussman and co-workers modified the design of
a drum and used a tapered, wheel-like disk as the collector.[64a] It was found that most of the fibers could be collected
on the sharp edge. The collected fibers were oriented parallel
1162

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 10. A) Schematic illustration of the setup used for electrospinning


nanofibers as uniaxially aligned arrays. The collector was constructed
from two conductive substrates separated by a void gap. B) The electrostatic force analysis of a charged nanofiber spanning across the gap. The
fiber experiences a strong stretching force (F) due to the Coulombic
interactions between the positive charges on the fiber and the negative,
image charges on the two grounded electrodes. C) Dark-field optical
micrograph of poly(vinyl pyrrolidone) (PVP) nanofibers collected across
the void gap between two silicon strips. D) SEM image of another sample taken from the region close to the edge of a void gap. As indicated by
arrows, some nanofibers near the gap took sharp turns to position them
perpendicular to the edge of the silicon strip. These figures were adapted
from [40b] with permission. Copyright American Chemical Society, 2003.

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

axially aligned over long length scales during the spinning


process (Fig. 10C). The introduction of an insulating gap to
the collector alters the configuration of electrostatic forces
acting on the fibers spanning across the gap. Under the action
of such electrostatic forces in opposite directions (Fig. 10B),
the charged fibers are stretched to align themselves perpendicular to each edge of the gap, even though abrupt changes
in the moving direction are required (Fig. 10D). The electrostatic repulsions between deposited nanofibers can further enhance the degree of alignment. Figure 11 shows SEM images
of various types of nanofibers that have been fabricated as
uniaxially aligned arrays using this simple setup. It is clear that
this new configuration is well-suited for use with materials
having a broad range of compositions and functionalities.
One of the remarkable features associated with this method
is that it is convenient to transfer the aligned fibers onto other
solid substrates for further processing steps and applications.
Single fibers could also be collected across the gap and transferred onto a substrate for the fabrication of single-fiberbased devices. In particular, it has been established that
uniaxially aligned fibers could be directly deposited on an
insulating substrate (e.g., quartz, PS) onto which the pair-wise
metal electrodes have been patterned.[68] In this variant, both
position and spatial orientation of the collected nanofibers
can be readily controlled by varying the layout or configuration of the electrodes. Furthermore, it was fairly easy to stack
aligned nanofibers into multilayered films by controlling the
configuration of electrodes and/or the scheme for applying
high voltage. Figure 12A shows a set of four gold electrodes
patterned on a quartz substrate. A cross-bar array of electrospun nanofibers (Fig. 12B) could be conveniently obtained by
sequentially grounding the electrode pairs 13 and 24. This
result demonstrated that the alignment and self-assembly of

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

Figure 12. A) Schematic illustration of a test pattern consisting of four


gold electrodes patterned on a quartz substrate. B) Optical micrograph
of a mesh made of poly(vinyl pyrrolidone) (PVP) nanofibers collected in
the central region of the four electrodes. During collection, the 13 and
24 electrode pairs were alternately grounded for ~ 5 s. The mean diameter of these fibers was around ~ 200 nm, as determined by SEM imaging.
This figure was adapted from [68].

Figure 11. SEM images of uniaxially aligned


nanofibers (AC) and nanotubes (D) made of
various materials: A) carbon; B) TiO2/PVP
composite; C) Sb-doped SnO2; and D) anatase. These figures were adapted from
[21b,40b] with permission. Copyright American
Chemical Society, 2003, 2004.

Adv. Mater. 2004, 16, No. 14, July 19

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1163

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

nanofibers into architectures useful for device fabrication


could be accomplished simultaneously in the production of
nanofibers to eliminate additional manipulation steps. As a
result, a variety of functional materials (including conjugated
organic polymers, carbon, ceramics, composites, and hollow
nanofibers) could all be fabricated as uniaxially aligned arrays
by collecting them with a pair of split electrodes. Such architectures should be immediately useful in the fabrication of
electronic and photonic devices.

5. Some Remarkable Features and Applications


5.1. Unique Features of Electrospun Nanofibers
Nanofibers exhibit a range of unique features and properties that distinguish themselves from 1D nanostructures fabricated using other techniques. For example, the electrospun
nanofiber is highly charged after it has been ejected from the
nozzle, and therefore it is possible to control its trajectory
electrostatically by applying an external electric field. As illustrated in Section 4.2, electrospun nanofibers could be readily
organized into a parallel array with the assistance of electrostatic interactions. This section highlights some other remarkable features in the context of potential applications.
5.1.1. Extremely Long Length
Compared with 1D nanostructures synthesized or fabricated using other chemical or physical methods, electrospun
nanofibers are extremely long.[3] Because electrospinning is a
continuous process, the fibers could be as long as several kilometers. This length scale is comparable to that of fibers manufactured by conventional drawing or spinning techniques. In
the electrospinning process, these long fibers can be assembled into a three-dimensional, nonwoven mat as a result
of bending instability of the spinning jet. Such a porous mat
can be immediately used for various applications. For instance, Pawlowski et al. demonstrated that lightweight wing
skins for a micro-air vehicle could be directly formed by electrospinning polymer nanofibers on a wing frame.[69] In addition to nonwoven mats, Xia and Li recently demonstrated that
individual fibers with several millimeters to centimeters long
could be manipulated individually using a collector containing
a void gap (e.g., metallic tweezers).[40b] The position and spatial orientation of an individual fiber can be easily controlled
by moving around the collector. Since the collector is a macroscopic object, the manipulation of individual nanofibers
could be achieved even without the assistance of a microscope.

1164

volume ratio. A high density of pores can also be formed as a


result of entanglement of nanofibers. Kim and co-workers
have evaluated the physical properties of nonwoven mats
composed of Nylon-6 nanofibers.[70] Depending on the fiber
diameter, the BrunauerEmmettTeller (BET) surface areas
of electrospun mats were between 9 and 51 m2 g1, while the
porosity varied from 25 % to 80 % and the pore size was in
the range of 2.737 to 0.167 lm. Although the specific surface
area of electrospun fibers is lower than mesoporous materials
such as molecular sieves, the pores in an electrospun, nonwoven mat are relatively large in size and all pores are fully interconnected to form a three-dimensional network. As a result,
the entire surface is fully accessible to chemical species. In
addition, as demonstrated in Section 3.3.3, the surface areas
of nanofibers could be further increased through the formation of much smaller pores in the surface of each individual
fiber by controlling the solutions and parameters for electrospinning.
5.1.3. Alignment on the Molecular Level
Electrospinning involves the rapid stretching of an electrified jet and rapid evaporation of the solvent. The polymer
chains are expected to experience an extremely strong shear
force during the electrospinning process. This shear force and
rapid solidification may prevent polymer chains from relaxing
back to their equilibrium conformations. As a result, the chain
conformation and crystallinity of the resultant polymer nanofibers should be different from products obtained by solutioncasting or conventional spinning processes. For example,
Foster and co-workers found that the polymer chains in
electrospun poly(ferrocenyldimethylsilane) were extended
and aligned along the fiber axis through electron diffraction
analysis.[71] Rabolt and co-workers have found that electrospinning Nylon-6 led to formation of the c-form, rather than
the a-form that was observed in films cast from solutions.[72]
Vancso and co-workers investigated the structures of electrospun PEO nanofibers by optical and atomic force microscopy
and concluded that the fibers possessed a surface layer, at
least, of highly ordered polymer chains.[73] Pedicini and Farris
characterized the stressstrain behavior of electrospun mats
of poly(urethane) (PU) fibers and found that the mats exhibited a fundamentally different stressstrain response curve in
uniaxial tensile tests.[74] This difference was believed to arise
from the orientation of chains in the electrospun fibers.
Martin and Lin studied the structures of electrospun fibers
composed of poly(hexyl isocynate), a liquid-crystalline polymer.[75] It was established that the orientation development in
the fibers was dependent on the diameter of the fibers.

5.1.2. High Surface Area and Complex Pore Structure

5.2. Applications of Electrospun Nanofibers

Compared with the fibers fabricated using a conventional


mechanical extrusion or spinning process, electrospun fibers
are much thinner in diameter and thus higher in surface-to-

As demonstrated in previous sections, electrospinning is a


remarkably simple and powerful technique for generating
ultrathin fibers from a wealth of different materials. The sim-

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

plicity of the fabrication scheme, the diversity of materials


suitable for use with electrospinning, as well as the unique
and interesting features associated with electrospun nanofibers, all make this technique and resultant structures attractive for a number of applications. Here we only discuss some
of these applications whose potentials have already been demonstrated to certain levels.
5.2.1. Nanofiber-Reinforced Composites
Fiber-based reinforcement represents one of the most effective strategies for enhancing the strength and other performance of a composite material. Because of the high surfaceto-volume ratio, use of nanofibers as the components may
significantly increase the interaction between the fibers and
the matrix material, leading to better reinforcement than conventional fibers. Reneker and Kim have demonstrated that
the mechanical strength of a rubber film could be significantly
improved by reinforcing with nanofibers electrospun from
poly(benzimidazole).[76] The Young's modulus measured for
the composite film was an order of magnitude greater and the
tear strength was twice as large as those of the pristine rubber
material. Vancso and Bergshoef have also shown that the
addition of electrospun Nylon nanofibers to an epoxy resin
could significantly improve both stiffness and mechanical
strength of the composite.[77] Because the cross-sections of
electrospun fibers are small, composites fabricated using this
method could maintain their optical transparency.
5.2.2. Nanofiber-Based Membranes and Smart Cloths
Gibson et al. have systematically investigated the transport
properties of porous membranes constructed from electrospun fibers.[78] Their results indicated that the electrospun
fiber mats had a higher convective resistance to air flow compared to normal clothing materials while the resistance to the
transport of water vapor was much lower than commercial
membrane laminates. In particular, the electrospun membranes exhibited the excellent ability to capture aerosol particles. The filtration efficiency was also high even though an
extremely thin layer of electrospun fibers was used. The lightweight, breathable mats made of electrospun nanofibers are
promising candidates for applications such as protective clothing. In fact, the high filtration efficiency of electrospun nonwoven membranes (especially for small particles) has long
been recognized by the filtration/separation industry for many
years.[79] Commercial filter products based on electrospun
fibers appeared more than twenty years ago. With the incorporation of magnetic components, it is expected that the separation efficiency for magnetic-active particles can be further
enhanced through the application of an external magnetic
field. In principle, it is possible to incorporate many types of
active components and species into electrospun nanofibers to
fabricate smart cloths that are responsive to all types of
changes in the surroundings.
Adv. Mater. 2004, 16, No. 14, July 19

5.2.3. Biomedical Usage

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

Because of their peculiar structures, electrospun nanofibers


have received much attention for use in tissue engineering.
Natural scaffolds for tissue growth are three-dimensional networks of nanometer-sized fibers made of various proteins.[80]
Successful tissue engineering requires man-made scaffolds to
bear similarity to the natural counterparts in terms of chemical
composition, morphology, and surface functional groups. Nonwoven mats of electrospun nanofibers are well known for their
interconnected, three-dimensional porous structures and relatively large surface areas. They provide a class of ideal materials to mimic the natural extra-cellular matrix required for tissue engineering. To this end, a number of biopolymers and
biodegradable polymers have been electrospun as nonwoven
membranes and their use as scaffolds for tissue engineering
have been demonstrated in recent years.[81] Several research
groups are also actively exploring other biomedical applications for electrospun nanofibers. For example, electrospun
fibers have been applied to wound dressing because porous
mats made of fibrous structures can protect the wound from
bacterial penetration via the aerosol particle capturing mechanism while providing a good path for the transport of vapor.[82]
In addition to the protection function, electrospun nonwoven
mats are also potentially useful as the supports or carriers for
drug delivery because of their high specific surface areas.[45]
5.2.4. Nanofiber-Based Supports for Enzymes and Catalysts
Nanofibers (both polymeric and ceramic) provide another
attractive class of solid supports for enzymes and conventional
catalysts thanks to their small sizes and large surface areas. To
this end, Wang and co-workers have demonstrated the use of
electrospun nanofibers as supports for enzymes.[83a] Bioactive
nanofibers have also been obtained by chemically attaching
enzyme to the surfaces of electrospun PS nanofibers with a
typical diameter of 120 nm. The as-prepared composites
showed high activities in both aqueous and organic media. Unlike solubilized or nanoparticle-supported enzymes, enzymecarrying fibers could be easily recovered from the reaction system. Hsieh and Xie also prepared enzyme-loaded fibers by
electrospinning solutions containing PEO or PVA and the
enzyme.[83b] The enzyme encapsulated in the electrospun fibers
was six times more reactive than that in a membrane film cast
from the same solution. Besides enzymes, nanofibers have
been explored as the solid supports for conventional catalysts.
For example, Erman and co-workers investigated the catalytic
activity of palladium-incorporated nanofibers for the selective
hydrogenation of dehydrolinalool and it was found that the
catalytic activity of such fibers was 4.5 times higher than the
traditional Pd catalyst supported on Al2O3 powders.[84]
5.2.5. Nanofiber-Based Sensors
Samuelson and co-workers have demonstrated the fabrication of fluorescence-based sensors with electrospun nano-

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1165

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

fibers containing a fluorescent indicator as the active component.[29] The sensitivities of such devices to metal ions (Fe3+ or
Hg2+) and 2,4-dinitrotoluene were two to three orders of magnitude greater than those based on continuous thin films. The
improvement was mainly caused by the availability of a larger
surface area and the relatively easy transport of analytes to
the active sites on each nanofiber. Gas sensors based on individual polyaniline nanofibers were recently fabricated by
Craighead and co-workers using their microfabricated electrospinning tip.[85] These authors also investigated the relationship between the response times and the diameters of
their nanofibers. Compared with film-based polyaniline sensors, both response time and sensitivity could be greatly improved. These demonstrations clearly indicate that electrospun nanofibers and their assemblies hold the promise as
active components for the sensing technology.
5.2.6. Nanofiber-Based Electrode Materials
Jo and co-workers have demonstrated that porous membranes made of electrospun fibers were good matrices for
holding polymer electrolytes and might find immediate use in
the fabrication of high-performance lithium batteries.[28a]
They have also found that the porous membrane (when
soaked with an electrolyte solution) exhibits an enhanced
ionic conductivity. In comparison, the polymer film made of
PVDF gel showed a relatively low value. This difference can
be attributed to the peculiar pore structure associated with
membranes composed of electrospun nanofibers. Both threedimensionally interconnected pore structure and the high
porosity associated with a mat of nanofibers facilitate the
transport of ions through the membrane. The high surface
area also improves the wettability of fibers by the electrolyte
solution, as well as the affinity between these two heterogeneous components. Yang and Kim have also evaluated the
performance of carbon nanofibers as an electrode material
for making supercapacitor.[86] The carbon nanofibers were
prepared by carbonization of electrospun fibers of PAN. They
observed a maximum capacitance of 175 F g1. By controlling
the activation temperature, the pore density, and structure
within the nanofibers could be varied to further improve the
performance of such a supercapacitor at higher current
densities.
5.2.7. Nanofiber-Based Electronic and Optical Devices
Like the semiconducting or metallic nanowires synthesized
using other methods, electrospun nanofibers with electrical
and electro-optical activities have also received much interest
in recent years because of their potential application in fabricating nanoscale electronic and optoelectronic devices. In this
regard, Wang and Santiago-Aviles have measured the electrical properties of partially graphitized carbon nanofibers prepared from electrospun nanofibers of PAN and a large negative magnetoresistance (75 %) was observed at 1.9 K.[87a]

1166

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

MacDiarmid and co-workers have electrospun nanofibers of


polyaniline/PEO blends with diameters smaller than 100 nm
and found that the conductivities of these fibers were dependent on their diameters.[87b] For fibers with diameters below
15 nm, they were electrically insulating. For asymmetric fibers
with one end thinner than the other end, their IV curves exhibited rectifying features. The same group has also demonstrated the fabrication of FETs from these fibers.[87c] In a different study, Xia and co-workers found that the intensity of
light scattered from nanofibers was dependent on the polarization direction of light.[40b] Based on this observation, arrays
of uniaxially aligned nanofibers have been demonstrated for
use as inexpensive optical polarizers.
5.2.8. Nanofibers as Sacrificial Templates
Like other 1D nanostructures, electrospun nanofibers can
also serve as sacrificial templates to generate 1D nanostructures with hollow interiors. A range of polymers, metals, and
ceramics have been prepared as nanotubes by coating electrospun polymer nanofibers with the particular material, followed by selective removal of the templates.[43a,51,60] Most
recently, Czaplewski et al. have demonstrated that electrospun fibers could be used as the templates to generate nanofluidic channels.[88a] In this case, nanofibers made of heat
depolymerizable polycarbonate were deposited on a target
substrate and then covered with spin-on glass. Nanofluidic
channels were obtained after the nanofibers had been selectively removed by heating. Unlike the channels that were fabricated by conventional lithographic techniques, the channels
obtained by templating against electrospun fibers exhibited
elliptical cross-sections and sharp corners were eliminated.
This structure may promote a smoother liquid flow through
the channels. In conjunction with contact photolithography,
they have also demonstrated the application of electrospun
fibers as the templates for fabricating nanoscale mechanical
oscillators.[88b] In these two applications, the size uniformity
and long axial length associated with electrospun fibers provide some immediate advantages over conventionally fabricated structures.

6. Concluding Remarks
The past few years have witnessed tremendous progress in
the area of electrospinning. Thanks to the efforts of many
research groups, electrospinning has now emerged as a technique capable of processing a rich variety of organic polymers,
ceramics, and composite materials into ultrathin fibers with
controllable diameters. The morphologies and internal structures of these fibers could also be tailored using a number of
physical and/or chemical methods. Furthermore, the setup for
electrospinning has been modified to directly generate nanofibers with coreshell or hollow structures, and as uniaxially
aligned arrays or layer-by-layer stacked films. In addition to
the advancement in experimental studies, theoretical model-

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

ing has also progressed to provide a better understanding on


the electrospinning process. All these research activities have
led to the exploitation of electrospun nanofibers in a broad
range of applications.
There is no doubt that electrospinning will become the technique of choice for generating 1D nanostructures in the future. However, many technical issues still need to be resolved
before this vision will come true. First of all, more experimental studies and theoretical modeling are required in order to
achieve a better control over the size and morphology of electrospun fibers. At the current stage of development, it is not
easy to electrospin fibers with diameters below 100 nm, in
particular, on the scale 1030 nm. Secondly, the diversity and
scope of materials that can be directly utilized with electrospinning have to be greatly expanded. Although more than 50
polymers with a range of functionalities have been electrospun as nanofibers, it remains a grand challenge to apply electrospinning to conjugated organic polymers as limited by their
molecular weights and/or solubilities. The important role
played by this class of polymers in molecular or flexible
electronics will certainly stimulate more interests and efforts
dedicated to the processing of these materials into fibrous
nanostructures. As shown in a recent demonstration,[89] these
polymers could be co-electrospun with other spinnable polymers using a spinneret containing two coaxial capillaries.
Thirdly, it is still necessary to systematically investigate the
correlation between the secondary structure of electrospun
nanofibers and the processing parameters (e.g., polymers or
solvents contained in the solution, electric field strength, and
the structure of spinneret). It is also important to organize
nanofibers of various types (e.g., porous, hollow, and core/
sheath) into well-defined arrays or hierarchical architectures
in order to fully achieve their potential in device fabrication
and materials science.
The capabilities of many well-established techniques for
materials processing can be greatly enhanced by combining
them with electrospinning. For instance, the solgel method
has been widely used for processing a rich variety of inorganic
materials into monoliths or thin films.[90] When combined with
electrospinning, it provides a remarkably simple and powerful
means for generating ceramic or composite nanofibers with
uniform diameters and low surface roughness. Such nanofibers might be useful as waveguiding structures. Furthermore,
ceramic materials play an important role in many other
technological areas such as catalysis, solar cells, rechargeable
batteries, supercapacitors, sensing, microelectromechanical
systems (MEMS), electronics, and photonics. Preparation of
ceramic materials in the form of nanofibers may significantly
improve their performances in the existing devices or open
the doors to new types of applications. For instance, the
research on engineering the secondary structures (porous,
core/sheath, and hollow) of electrospun nanofibers will provide a new platform for designing advanced electrode materials, catalyst supports, and sensing devices. In particular, hollow nanofibers with circular cross-sections are ideal channels
for fabricating nanofluidic devices. They can also serve as

Adv. Mater. 2004, 16, No. 14, July 19

templates for generating new 1D nanostructures.


Besides their direct use as active components, electrospun
nanofibers may also find use in a number of nanoscale fabrication or manipulation tasks. For example, electrospun nanofibers are very long and it is much easier to manipulate individual electrospun nanofibers than nanowires or nanorods
synthesized using other methods. When a pair of split electrodes are used as the collector, the nanofibers can be directly
deposited across the two electrodes to form a natural connection between the nano- and macroscopic world. In conjunction with conventional fabrication techniques (e.g., used as
masks for etching, deposition, or exposure), electrospun nanofibers may provide an alternative route to functional nanostructures, devices, and systems. Furthermore, as demonstrated in Section 3.2.3, nanostructures of various morphologies
and compositions could be readily incorporated into the interiors of nanofibers. As a result, nanofibers and their 2D or 3D
assemblies could serve as a new platform to manipulate the
spatial locations and orientations of much smaller features.
For example, nanoscale objects encapsulated in an electrospun nanofiber could be placed at the desired location on a
solid substrate by moving the nanofiber with a longer axial
dimension.
It is expected that research on electrospinning will become
more interdisciplinary in the near future. Traditionally, this
field has been dominated by people with expertise in areas
such as polymer and textile engineering. As the capability of
this technique is extended to functional materials other than
organic polymers and more new applications are demonstrated, further development of this technique will need extensive
collaborations between researchers working in different fields.
With the involvement of a bigger scientific and engineering
community, electrospinning will surely become one of the
most powerful tools for fabricating nanostructures and nanomaterials with the broadest range of functionalities and applications.

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

Received: May 11, 2004

[1]

[2]
[3]

[4]

[5]

[6]

http://www.advmat.de

Recent reviews: a) Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers,


B. Gates, Y. Yin, F. Kim, H. Yan, Adv. Mater. 2003, 15, 353. b) J. Hu,
T. W. Odom, C. M. Lieber, Acc. Chem. Res. 1999, 32, 435.
See a recent special issue in Adv. Mater. 2003, 15, pp. 351456.
a) D. H. Reneker, I. Chun, Nanotechnology 1996, 7, 216. b) A. Frenot, I. S. Chronakis, Curr. Opin. Colloid Interface Sci. 2003, 8, 64.
c) Z.-M. Huang, Y.-Z. Zhang, M. Kotaki, S. Ramakrishna, Compos.
Sci. Technol. 2003, 63, 2223.
a) D. H. Reneker, A. L. Yarin, H. Fong, S. Koombhongse, J. Appl.
Phys. 2000, 87, 4531. b) Y. M. Shin, M. M. Hohman, M. P. Brenner,
G. C. Rutledge, Appl. Phys. Lett. 2001, 78, 1149.
a) A. G. Bailey, Electrostatic Spraying of Liquids, Wiley, New York
1988. b) J. B. Fenn, M. Mann, C. K. Meng, S. F. Wong, C. M. Whitehouse, Science 1989, 246, 64. c) I. G. Loscertales, A. Barrero,
I. Guerrero, R. Cortijo, M. Marquez, A. M. Gan-Calvo, Science
2002, 295, 1695. d) P. J. Hull, J. L. Hutchison, O. V. Salata, P. J. Dobson, Adv. Mater. 1997, 9, 413.
A. Formalas, US patent 1 975 504, 1934.

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1167

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

1168

[7] a) P. K. Baumgarten, J. Colloid Interface Sci. 1971, 36, 71. b) L. Larrondo, R. Manley, J. Polym. Sci: Polym. Phys. Ed. 1981, 19, 909.
c) L. Larrondo, R. Manley, J. Polym. Sci: Polym. Phys. Ed. 1981, 19,
921. d) L. Larrondo, R. Manley, J. Polym. Sci: Polym. Phys. Ed.
1981, 19, 933.
[8] J. Doshi, D. H. Reneker, J. Electrost. 1995, 35, 151.
[9] R. Kessick, J. Fenn, G. Tepper, Polymer 2004, 45, 2981.
[10] a) A. L. Yarin, S. Koombhongse, D. H. Reneker, J. Appl. Phys. 2001,
90, 4836. b) Y. M. Shin, M. M. Hohman, M. P. Brenner, G. C. Rutledge, Polymer 2001, 42, 9955.
[11] a) M. M. Hohman, M. Shin, G. C. Rutledge, M. P. Brenner, Phys.
Fluidics 2001, 13, 2201. b) M. M. Hohman, M. Shin, G. C. Rutledge,
M. P. Brenner, Phys. Fluidics 2001, 13, 2221. c) S. V. Fridrikh, J. H.
Yu, M. P. Brenner, G. C. Rutledge, Phys. Rev. Lett. 2003, 90, 144 502.
[12] a) J. J. Feng, Phys. Fluidics 2002, 14, 3912. b) J. J. Feng, J. Non-Newtonian Fluid Mech. 2003, 116, 55.
[13] J.-S. Kim, D. H. Reneker, Polym. Eng. Sci. 1999, 39, 849.
[14] D. Fang, B. S. Hsiao, B. Chu, Polym. Prepr. (Am. Chem. Soc., Div.
Polym. Chem.) 2003, 44, 59.
[15] G. M. Whitesides, P. E. Laibinis, Langmuir 1990, 6, 87.
[16] W. Lang, Mater. Sci. Eng., R 1996, 17, 1.
[17] http://www.polymicro.com/
[18] a) J. Kameoka, H. G. Craighead, Appl. Phys. Lett. 2003, 83, 371.
b) J. Kameoka, R. Orth, Y. Yang, D. Czaplewski, R. Mathers, G. W.
Coates, H. G. Craighead, Nanotechnology 2003, 14, 1124.
[19] R. Kessick, G. Tepper, Appl. Phys. Lett. 2003, 83, 557.
[20] a) P. Gupta, G. L. Wilkes, Polymer 2003, 44, 6353. b) S. Madhugiri,
A. Dalton, J. Gutierrez, J. P. Ferraris, K. J. Balkus, Jr., J. Am. Chem.
Soc. 2003, 125, 14 531.
[21] a) Z. Sun, E. Zussman, A. L. Yarin, J. H. Wendorff, A. Greiner,
Adv. Mater. 2003, 15, 1929. b) D. Li, Y. Xia, Nano Lett. 2004, 4, 933.
c) I. G. Loscertales, A. Barrero, M. Marquez, R. Spretz, R. VelardeOrtiz, G. Larsen, J. Am. Chem. Soc. 2004, 126, 5376.
[22] G. Larsen, R. Spretz, R. Velarde-Ortiz, Adv. Mater. 2004, 16, 166.
[23] a) H. Fong, I. Chun, D. H. Reneker, Polymer 1999, 40, 4585.
b) K. H. Lee, K. Y. Kim, H. J. Bang, Y. H. Jung, S. G. Lee, Polymer
2003, 44, 4029. c) J. M. Deitzel, K. Kleinmeyer, D. Harris, N. C. B.
Tan, Polymer 2001, 42, 261. d) H. Fong, D. H. Reneker, J. Polym.
Sci: Polym. Phys. Ed. 1999, 37, 3488.
[24] a) L. Rayleigh, Proc. R. Soc. London 1882, 14, 184. b) J. N. Smith,
R. C. Flagan, J. L. Beauchamp, J. Phys. Chem. A 2002, 106, 9957.
[25] a) S. Koombhongse, W. Liu, D. H. Rekener, J. Polym. Sci: Polym.
Phys. Ed. 2001, 39, 2598. b) H. Fong, W. Liu, C.-S. Wang, R. A. Vaia,
Polymer 2002, 43, 775. c) R. V. N. Krishnappa, K. Desai, C. Sung,
J. Mater. Sci. 2003, 38, 2357. d) W. D. Bates, C. P. Barnes, Z. Ounaies,
G. E. Wnek, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.)
2003, 44, 114.
[26] For example, a) M. G. McKee, G. L. Wilkes, R. H. Colby, T. E.
Long, Macromolecules 2004, 37, 1760. b) S. A. Theron, E. Zussman,
A. L. Yarin, Polymer 2004, 45, 2017. c) W. K. Son, J. H. Youk, T. S.
Lee, W. H. Park, Polymer 2004, 45, 2959. d) K. H. Lee, H. Y. Kim,
M. S. Khil, Y. M. Ra, D. R. Lee, Polymer 2003, 44, 1287. e) Z. Jun,
H. Hou, A. Schaper, J. H. Wendorff, A. Greiner, e-Polymers 2003,
no. 1. f) L. Yan, T. W. Haas, A. Guiseppi-Elie, G. L. Bowlin, D. G.
Simpson, G. E. Wnek, Chem. Mater. 2003, 15, 1860. g) D. H. Reneker, W. Kataphinan, A. Theron, E. Zussman, A. L. Yarin, Polymer
2002, 43, 6785. h) M. M. Demir, I. Yilgor, E. Yilgor, B. Erman, Polymer 2002, 43, 3303. i) X. Zong, K. Kim, D. Fang, S. Ran, B. S. Hsiao,
B. Chu, Polymer 2002, 43, 4403. j) C. J. Buchko, L. C. Chen, Y. Shen,
D. C. Martin, Polymer 1999, 40, 7397.
[27] A. G. MacDiarmid, W. E. Jones, Jr., I. D. Norris, J. Gao, A. T. Johnson, Jr., N. J. Pinto, J. Hone, B. Han, F. K. Ko, H. Okuzaki, M. Llaguno, Synth. Met. 2001, 119, 27.
[28] a) S. W. Choi, S. M. Jo, W. S. Lee, Y.-R. Kim, Adv. Mater. 2003, 15,
2027. b) W. D. Bates, C. P. Barnes, Z. Ounaies, G. E. Wnek, Polym.
Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 2003, 44, 114.

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

[29] X. Wang, C. Drew, S.-H. Lee, K. J. Senecal, J. Kumar, L. A. Samuelson, Nano Lett. 2002, 2, 1273.
[30] W. R. Caseri, H. D. Chanzy, K. Feldman, M. Fontana, P. Smith, T. A.
Tervoort, J. G. P. Goossens, E. W. Meijer, W. Egbert, A. P. H. J.
Schenning, I. P. Dolbnya, M. G. Debije, M. P. de Haas, J. M. Warman, A. M. van de Craats, R. H. Friend, H. Sirringhaus, N. Stutzmann, Adv. Mater. 2003, 15, 125.
[31] a) E. D. Boland, G. E. Wnek, D. G. Simpson, K. J. Pawlowski, G. L.
Bowlin, J. Macromol. Sci., Pure Appl. Chem. 2001, A38, 1231.
b) J. Zeng, X. Chen, X. Xu, Q. Liang, X. Bian, L. Yang, X. Jing,
J. Appl. Polym. Sci. 2003, 89, 1085. c) L. Huang, R. A. McMillan,
R. P. Apkarian, B. Pourdeyhimi, V. P. Conticello, E. L. Chaikof,
Macromolecules 2000, 33, 2989. d) E.-R. Kenawy, J. M. Layman,
J. R. Watkins, G. L. Bowlin, J. A. Matthews, D. G. Simpson, G. E.
Wnek, Biomaterials 2003, 24, 907.
[32] a) J. A. Matthews, G. E. Wnek, D. G. Simpson, G. L. Bowlin, Biomacromolecules 2002, 3, 232. b) G. E. Wnek, C. E. Marcus, D. G.
Simpson, G. L. Bowlin, Nano Lett. 2003, 3, 213. c) S. Sukigara,
M. Gandhi, J. Ayutsede, M. Micklus, F. Ko, Polymer 2003, 44, 5721.
d) K. Ohgo, C. Zhao, M. Kobayashi, T. Asakura, Polymer 2003, 44,
841. e) H. Jiang, D. Fang, B. S. Hsiao, B. Chu, W. Chen, Biomacromolecules 2004, 5, 326. f) S.-W. Lee, A. M. Belcher, Nano Lett. 2004,
4, 387. g) X. Fang, D. H. Reneker, J. Macromol. Sci. Phys. 1997, B36,
169.
[33] H.-J. Jin, S. V. Fridrikh, G. C. Rutledge, D. L. Kaplan, Biomacromolecules 2002, 3, 1233.
[34] a) I. D. Norris, M. M. Shaker, F. K. Ko, A. G. MacDiarmid, Synth.
Met. 2000, 114, 109. b) K. Desai, C. Sung, Mater. Res. Soc. Symp.
Proc. 2003, 788, 209. c) P. K. Kahol, N. J. Pinto, Synth. Met. 2004,
140, 269.
[35] K. Kim, M. Yu, X. Zong, J. Chiu, D. Fang, Y.-S. Seo, B. S. Hsiao,
B. Chu, M. Hadjiargyrou, Biomaterials 2003, 24, 4977.
[36] G. Larsen, R. Velarde-Ortiz, K. Minchow, A. Barrero, I. G. Loscertales, J. Am. Chem. Soc. 2003, 125, 1154.
[37] a) Y. Wang, R. Furlan, I. Ramos, J. I. Santigano-Aviles, Appl. Phys.
A 2004, 78, 1043. b) Y. Wang, J. J. Santigano-Aviles, Nanotechnology
2004, 15, 32.
[38] a) S.-S. Choi, S. G. Lee, S. S. Im, S. H. Kim, J. Mater. Sci. Lett. 2003,
22, 891. b) W. Kataphinan, R. Teye-Mensah, E. A. Evans, R. D.
Ramsier, D. H. Reneker, D. J. Smith, J. Vac. Sci. Technol., A 2003,
21, 1574.
[39] a) S. Madhugiri, W. Zhou, J. P. Ferraris, K. J. Balkus, Jr., Microporous Mesoporous Mater. 2003, 63, 75. b) S. Madhugiri, A. Chacko,
J. P. Ferraris, K. J. Balkus, Jr., Polym. Prepr. (Am. Chem. Soc., Div.
Polym. Chem.) 2003, 44, 86.
[40] a) D. Li, Y. Xia, Nano Lett. 2003, 3, 555. b) D. Li, Y. Wang, Y. Xia,
Nano Lett. 2003, 3, 1167. c) D. Li, T. Herricks, Y. Xia, Appl. Phys.
Lett. 2003, 83, 4586.
[41] a) H. Dai, J. Gong, H. Kim, D. Lee, Nanotechnology 2002, 13, 674.
b) P. Viswanathamurthi, N. Bhattarai, H. Y. Kim. D. R. Lee, S. R.
Kim, M. A. Morris, Chem. Phys. Lett. 2003, 374, 79. c) C. Shao,
H.-Y. Kim, J. Gong, B. Ding, D.-R. Lee, S.-J. Park, Mater. Lett. 2003,
57, 1579. d) H. Guan, C. Shao, B. Chen, J. Gong, X. Yang, Inorg.
Chem. Commun. 2003, 6, 1409. e) H. Guan, C. Shao, B. Chen,
J. Gong, X. Yang, Inorg. Chem. Commu. 2003, 6, 1302. f) X. Yang,
C. Shao, H. Guan, X. Li, J. Gong, Inorg. Chem. Commun. 2004, 7,
176. g) B. Ding, H. Kim, C. Kim, M. Khil, S. Park, Nanotechnology
2003, 14, 532. h) P. Viswanathamurthi, N. Bhattarai, H. Y.
Kim. D. R. Lee, Scr. Mater. 2003, 49, 577. i) H. Guan, C. Shao,
S. Wen. B. Chen, J. Gong, X. Yang, Mater. Chem. Phys. 2003, 6, 1302.
j) P. I. Gouma, Rev. Adv. Mater. Sci. 2003, 5, 147. k) N. Dharmaraj,
H. C. Park, B. M. Lee, P. Viswanathamurthi, H. Y. Kim, D. R. Lee,
Inorg. Chem. Commun. 2004, 7, 431.
[42] a) K. J. Senecal, D. P. Ziegler, J. He, R. Mosurkal, H. Schreuder-Gibson, L. A. Samuelson, Mater. Res. Soc. Symp. Proc. 2002, 788, 285.
b) A. Pedicini, R. J. Farris, J. Polym. Sci., Polym. Phys. Ed. 2004, 42,

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

[43]

[44]

[45]

[46]
[47]

[48]

[49]
[50]

[51]
[52]
[53]

[54]
[55]

[56]
[57]
[58]

752. c) N. Viriyabanthorn, J. L. Mead, R. G. Stacer, C. Sung, Polym.


Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 2003, 44, 136. d) Q. B.
Yang, D. M. Li, Y. L. Hong, Z. Y. Li, C. Wang, S. L. Qiu, Y. Wei,
Synth. Met. 2003, 137, 973. e) H. Yang, L. Loh, T. Han, F. Ko, Polym.
Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 2003, 44, 163.
a) H. Hou, Z. Jun, A. Reuning, A. Schaper, J. H. Wendorff, A. Greiner, Macromolecules 2002, 16, 2429. b) H. Hou, D. H. Reneker, Adv.
Mater. 2004, 16, 69. c) X. Hao, Z. Li, C. Wang, Polym. Prepr. (Am.
Chem. Soc., Div. Polym. Chem.) 2003, 44, 149.
a) F. Ko, Y. Gogotsi, A. Ali, N. Naguib, H. Ye, G. Yang, C. Li,
P. Willis, Adv. Mater. 2003, 15, 1161. b) Y. Dror, W. Salalha, R. L.
Khalfin, Y. Cohen, A. L. Yarin, E. Zussman, Langmuir 2003, 19,
7012. c) C. Seoul, Y.-T. Kim, C.-K. Baek, J. Polym. Sci., Polym. Phys.
Ed. 2003, 41, 1572. d) J. A. Stuckey, M. D. Alexander, Jr., B. M.
Black, J. D. Henes, Polym. Prepr. (Am. Chem. Soc., Div. Polym.
Chem.) 2003, 44, 141. e) R. Sen, B. Zhao, D. Perea, M. E. Itkis,
H. Hu, J. Love, E. Bekyarova, R. C. Haddon, Nano Lett. 2004, 4,
459.
a) Y. K. Luu, K. Kim, B. S. Hsiao, B. Chu, M. Hadjiargyrou, J. Controlled Release 2003, 89, 341. b) J. Zeng, X. Xu, X. Chen, Q. Liang,
X. Bian, L. Yang, X. Jing, J. Controlled Release 2003, 92, 227.
E. H. Sanders, R. Kloefkorn, G. L. Bowlin, D. G. Simpson, G. E.
Wnek, Macromolecules 2003, 36, 3803.
a) K. Kim, C. Chang, X. Zong, D. Fang, B. S. Hsiao, B. Chu, M. Hadjiargyrou, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 2003,
44, 98. b) C. Drew, X. Wang, L. A. Samuelson, J. Kumar, J. Macromol. Sci., Pure Appl. Chem. 2003, A40, 1415.
a) X. Wang, Y.-G. Kim, C. Drew, B.-C. Ku, J. Kumar, L. A. Samuelson, Nano Lett. 2004, 4, 331. b) C. Drew, X. Liu, S. Ziegler, X. Wang,
F. F. Bruno, J. Whitten, L. A. Samuelson, J. Kumar, Nano Lett. 2003,
3, 143.
N. J. Pinto, P. Carrin, J. X. Quiones, Mater. Sci. Eng., A 2004, 366,
1.
a) M. Bognitzki, Z. Jia, A. K. Schaper, R. B. Wehrspohn, U. Gsele,
J. H. Wendorff, Adv. Mater. 2000, 12, 637. b) W. Liu, M. Graham,
E. A. Evans, D. H. Reneker, J. Mater. Res. 2002, 17, 2002.
D. Li, J. McCann, M. Gratt, Y. Xia, Chem. Phys. Lett. 2004, submitted.
X. Zong, S. Ran, D. Fang, B. S. Hsiao, B. Chu, Polymer 2003, 44,
4959.
a) I. Chun, D. H. Reneker, H. Fong, X. Fang, J. Deitzel, N. B. Tan,
K. Kearns, J. Adv. Mater. 2003, 31, 37. b) S. H. Park, C. Kim, Y. O.
Choi, K. S. Yang, Carbon 2003, 41, 2655. c) K. S. Yang, D. D. Edie,
D. Y. Lim, Y. M. Kim, Y. O. Choi, Carbon 2003, 41, 2039.
d) Y. Wang, S. Serrano, J. J. Santigano-Aviles, Synth. Met. 2003, 138,
423. e) D. L. Gee, G. E. Wnek, S. M. Zhuang, J. M. Layman, P. Lipowiez, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 2003,
44, 120.
H. Liu, Y.-L. Hsieh, J. Polym. Sci., Polym. Phys. Ed. 2003, 41, 953.
a) K. Nagapudi, W. T. Brinkman, J. E. Leisen, L. Huang, R. A.
McMillan, R. P. Apkarian, V. P. Conticello, E. L. Chaikof, Macromolecules 2002, 35, 1730. b) J. Zeng, H. Hou, J. H. Wendorff,
A. Greiner, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.)
2003, 44, 174. c) D. A. Baker, P. J. Brown, Polym. Prepr. (Am. Chem.
Soc., Div. Polym. Chem.) 2003, 44, 118.
E.-R. Kenawy, Y. R. Abdel-Fattah, Macromol. Biosci. 2002, 2, 261.
M. Wei, J. Mead, C. Sung, Polym. Prepr. (Am. Chem. Soc., Div.
Polym. Chem.) 2003, 44, 79.
a) R. H. Baughman, A. A. Zakhidov, W. A. de Heer, Science 2002,
297, 787. b) R. Tenne, Angew. Chem. Int. Ed. 2003, 42, 5124. c) C. R.
Martin, Science 1994, 266, 1961. d) J. Goldberger, R. He, Y. Zhang,
S. Lee, H. Yan, H.-J. Choi, P. Yang, Nature 2003, 422, 599.
e) B. Mayers, X. Jiang, D. Sunderland, B. Cattle, Y. Xia, J. Am.
Chem. Soc. 2003, 125, 13 364. f) M. Steinhart, Z. Jia, A. K. Schaper,
R. B. Wehrspohn, U. Gsele, J. H. Wendorff, Adv. Mater. 2003, 15,
706. g) Y. Sun, Y. Xia, Adv. Mater. 2004, 16, 264.

Adv. Mater. 2004, 16, No. 14, July 19

[59] a) R. A. Caruso, J. H. Schattka, A. Greiner, Adv. Mater. 2001, 13,


1577. b) H. Dong, S. Prasad, V. Nyame, W. E. Jones, Jr., Chem.
Mater. 2004, 16, 371. c) F. Ochanda, A. Atkinson, W. E. Jones, Jr.,
Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 2003, 44, 161.
[60] M. Bognitzki, T. Frese, M. Steinhart, A. Greiner, J. H. Wendorff,
A. Schaper, M. Hellwig, Polym. Eng. Sci. 2001, 41, 982.
[61] M. Bognitzki, W. Czado, T. Frese, A. Schaper, M. Hellwig, M. Steinhart, A. Greiner, J. H. Wendorff, Adv. Mater. 2001, 13, 70.
[62] a) S. Megelski, J. S. Stephens, D. B. Chase, J. F. Rabolt, Macromolecules 2002, 35, 8456. b) C. L. Casper, J. S. Stephens, N. G. Tassi,
D. B. Chase, J. F. Rabolt, Macromolecules 2004, 37, 573.
[63] a) N. I. Kovtyukhova, T. E. Mallouk, Chem. Eur. J. 2002, 8, 4354.
b) Y. Huang, X. Duan, Q. Wei, C. M. Lieber, Science 2001, 291, 630.
c) F. Favier, E. C. Walter, M. P. Zach, T. Benter, R. M. Penner,
Science 2001, 293, 2227. d) N. A. Melosh, A. Boukai, F. Diana,
B. Gerardot, A. Badolato, P. M. Petroff, J. R. Heath, Science 2003,
300, 112. e) F. Kim, S. Kwan, J. Akana, P. Yang, J. Am. Chem. Soc.
2001, 123, 4360.
[64] a) A. Theron, E. Zussman, A. L. Yarin, Nanotechnology 2001, 12,
384. b) E. Zussman, A. Theron, A. L. Yarin, Appl. Phys. Lett. 2003,
82, 973.
[65] a) B. Sundaray, V. Subramanian, T. S. Natarajan, Appl. Phys. Lett.
2003, 82, 973. b) C. Y. Xu, R. Inai, M. Kotaki, S. Ramakrishna, Biomaterials 2004, 25, 877.
[66] a) H. Fong, W. D. Liu, C. S. Wang, R. A. Vaia, Polymer 2002, 43,
775. b) R. Dersch, T. Liu, A. K. Schaper, A. Greiner, J. H. Wendorff,
J. Polym. Sci. A 2003, 41, 545.
[67] J. M. Deitzel, J. D. Kleinmeyer, J. K. Hirvonen, N. C. B. Tan, Polymer 2001, 42, 8163.
[68] D. Li, Y. Wang, Y. Xia, Adv. Mater. 2004, 16, 361.
[69] K. J. Pawlowski, H. L. Belvin, D. L. Raney, J. Su, J. S. Harrison, E. J.
Siochi, Polymer 2003, 44, 1309.
[70] Y. J. Ryu, H. Y. Kim, K. H. Lee, H. C. Park, D. R. Lee, Eur. Polym.
J. 2003, 39, 1883.
[71] Z. Chen, M. D. Foster, W. Zhou, H. Fong, D. H. Reneker, Macromolecules 2001, 34, 6156.
[72] J. S. Stephens, D. B. Chase, J. F. Rabolt, Macromolecules 2004, 37, 877.
[73] R. Jaeger, H. Schnherr, G. J. Vancso, Macromolecules 1996, 29, 7634.
[74] A. Pedicini, R. J. Farris, Polymer 2003, 44, 6857.
[75] D. Y. Lin, D. C. Martin, Polym. Prepr. (Am. Chem. Soc., Div. Polym.
Chem.) 2003, 44, 70.
[76] J.-S. Kim, D. H. Reneker, Polym. Compos. 1999, 20, 124.
[77] M. M. Bergshoef, G. J. Vancso, Adv. Mater. 1999, 11, 1362.
[78] P. Gibson, H. Schreuder-Gibson, D. Rivin, Colloids Surf. A 2001,
187188, 469.
[79] http://www.donaldson.com/
[80] K. E. Kadler, D. F. Holmes, J. A. Trotter, J. A. Chapman, Biochem.
J. 1996, 316, 1.
[81] a) W.-J. Li, C. T. Laurencin, E. J. Caterson, R. S. Tuan, F. K. Ko,
J. Bio. Mater. Res. 2002, 60, 613. b) Y. Yoshimoto, Y. M. Shin, H. Terai, J. P. Vacanti, Biomaterials 2003, 24, 2077. c) X. M. Mo, C. Y. Xu,
M. Kotaki, S. Ramakrishna, Biomaterials 2004, 25, 1883. d) H.-J. Jin,
J. Chen, V. Karageorgiou, G. H. Altman, D. L. Kaplan, Biomaterials
2004, 25, 1039.
[82] a) M. Khil, D. Cha, H. Kim, I. Lim, N. Bhattarai, J. Biomed.
Mater. Res., Part B 2003, 67B, 675. b) X. Zong, K. Kim, J. Chiu,
B. S. Hsiao, B. Chu, S. Li, B. Garlick, C. Brathwaite, T. Zimmerman, D. Fang, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.)
2003, 44, 89. c) J. M. Layman, E.-R. Kenawy, J. R. Watkins,
M. E Carr, Jr. G. Bowlin, G. E. Wnek, Polym. Prepr. (Am. Chem.
Soc., Div. Polym. Chem.) 2003, 44, 94. d) B.-M. Min, G. Lee, S. H.
Kim, Y. S. Nam, T. S. Lee, W. H. Park, Biomaterials 2004, 25,
1289.
[83] a) H. Jia, G. Zhu, B. Vugrinovich, W. Kataphinan, D. H. Reneker,
P. Wang, Biotechnol. Prog. 2002, 18, 1027. b) J. Xie, Y.-L. Hsieh,
J. Mater. Sci. 2003, 38, 2125.

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

1169

REVIEW

D. Li, Y. Xia/Electrospinning of Nanofibers

[84] M. M. Demir, M. A. Gulgun, Y. Z. Menceloglu, B. Erman, S. S.


Abramchuk, E. E. Makhaeva, A. R. Khokhlov, V. G. Matveeva,
M. G. Sulman, Macromolecules 2004, 37, 573.
[85] H. Liu, J. Kameoka, D. A. Czaplewski, H. G. Craighead, Nano Lett.
2004, 4, 671.
[86] C. Kim, K. S. Yang, Appl. Phys. Lett. 2003, 83, 1216.
[87] a) Y. Wang, J. J. Santigano-Aviles, J. Appl. Phys. 2003, 94, 1721.
b) Y. Zhou, M. Freitag, J. Hone, C. Staii, A. T. Johnson, Jr., N. J.
Pinto, A. G. MacDiarmid, Appl. Phys. Lett. 2003, 83, 3800. c) N. J.
Pinto, A. T. Johnson, Jr., A. G. MacDiarmid, C. H. Mueller,
N. Theofylaktos, D. C. Robinson, F. A. Miranda, Appl. Phys. Lett.
2003, 83, 4244.

[88] a) D. A. Czaplewski, J. Kameoka, R. Mathers, G. W. Coates, H. G.


Craighead, Appl. Phys. Lett. 2003, 83, 4836. b) D. A. Czaplewski,
S. S. Verbridge, J. Kameoka, H. G. Craighead, Nano Lett. 2004, 4,
437.
[89] D. Li, A. Babel, S. A. Jenekhe, Y. Xia, Adv. Mater. 2004, submitted.
[90] a) J. D. Wright, N. A. J. M. Sommerdijk, Sol-Gel Materials: Chemistry and Applications, Gordon and Breach Science, Amsterdam 2001.
b) C. Sanchez, B. Lebeau, F. Chaput, J.-P. Boilot, Adv. Mater. 2003,
15, 1969.

______________________

1170

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

You might also like