Sie sind auf Seite 1von 16

Composites Science and Technology 70 (2010) 703718

Contents lists available at ScienceDirect

Composites Science and Technology


journal homepage: www.elsevier.com/locate/compscitech

Review

Electrospinning of polymer nanobers: Effects on oriented morphology, structures and tensile properties
Avinash Baji a, Yiu-Wing Mai a,b,*, Shing-Chung Wong c, Mojtaba Abtahi a, Pei Chen c
a

Centre for Advanced Materials Technology (CAMT), School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, Sydney, NSW 2006, Australia Department of Mechanical Engineering, Hong Kong Polytechnic University, Kowloon, Hong Kong, China c Department of Mechanical Engineering, The University of Akron, Akron, Ohio 44325, USA
b

a r t i c l e

i n f o

a b s t r a c t
The interest in fabrication of nanobers using electrospinning method has attracted considerable attention due to its versatile maneuverability of producing controlled ber structures, porosity, orientations and dimensions. Although the process appears to be simple and straightforward, an understanding of the technique and its inuence on the morphology, structural and mechanical properties is still not completely clear. Recently, the size effect on the mechanical properties was reported for bers across different length scales. Both modulus and strength of poly(e-capro-lactone) (PCL) bers were found to increase signicantly when the diameter of the bers was reduced to below $500 nm. In this article, for the rst time, we critically review and evaluate the role of the microstructures on the ber deformation behavior and present possible explanations for the enhanced properties of the nanobers. Our discussions are focused on the techniques to obtain controlled structures and the mechanisms behind the size effect in electronspun bers are given. In-depth understanding of these mechanisms can provide fruitful outcomes in the development of advanced nanomaterials for devices and miniaturized load-bearing applications. 2010 Elsevier Ltd. All rights reserved.

Article history: Received 19 October 2009 Received in revised form 12 January 2010 Accepted 14 January 2010 Available online 20 January 2010 Keywords: A. Fibers A. Nano composites B. Mechanical properties D. X-ray diffraction (XRD) E. Electro-spinning

Contents 1. 2. 3. 4. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Electrospinning theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control of fiber diameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alignment of fibers and fiber collection methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Rotating drum collector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Rotating disk collector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Static parallel electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structural properties of electrospun fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Molecular orientation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Crystallinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Effect of fiber diameter on structural properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Effect of collector on the structural properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mechanical properties of the fibers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1. Effect of structural morphology on tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2. Effect of collector type on tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1. Stationary collector. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.2. Rotational collector. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3. Effect of fiber diameter on tensile properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Prospective applications of electrospun fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1. Fiber composites for tissue engineering applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704 705 706 706 707 708 708 709 709 710 710 710 711 711 712 712 712 712 713 713

5.

6.

7.

* Corresponding author. Address: Centre for Advanced Materials Technology (CAMT), School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, Sydney, NSW 2006, Australia. Tel.: +61 2 9351 2290; fax: +61 2 9351 3760. E-mail address: y.mai@usyd.edu.au (Y.-W. Mai). 0266-3538/$ - see front matter 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.compscitech.2010.01.010

704

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

8.

7.2. Electrospun fiber reinforced composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3. Conductive fiber composites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4. Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5. Filler reinforced fiber systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Concluding remarks and future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

714 714 715 715 716 716 716

1. Introduction The drive for ultra-lightweight yet strong structures for devices and miniaturized applications has motivated novel designs using polymer nanobers [1]. Electrospinning has emerged as a powerful technique for producing high strength bers due to its versatility, ease of use, ability to align structures and control ber diameters [29]. Some of these unique features cannot be otherwise achieved by conventional ber processing techniques. Another merit is that under the inuence of an electric eld, electrospinning self-assembles dispersed llers along the axial direction such that composites can be formed by imposing additional spatial connement to the polymer chains [5,10,11]. These reinforced bers display superior properties and function as basic building blocks for the fabrication of high strength structures using a bottom-up approach. For example, carbon nanotubes (CNTs) and carbon black (CB) particles are among the commonly used llers which are dispersed within the bers to mimic the functionality of silk bers for high strength and toughness applications [1]. This feature of dispersing ller materials can be easily extended to other applications such as ltration [12,13], tissue engineering [9,14,15], precursor for fabricating nanober composites [1618] and advanced nanomaterials [4,19] etc. Despite possessing these unique features, one of the main challenges in this area is to characterize the tensile behavior of the nanobers. This could be due to the difculty in handling the nanobers and also due to the low load required for the deformation. Hence, in most cases, the mechanical integrity of the bers and ber network structures is least understood and an understanding of the phenomenon is urgently needed. Few researchers actively pursued to characterize the mechanical deformation characteristics of the bers by recording the stressstrain behavior of

the electrospun non-woven fabrics. However, this method cannot be deemed suitable because the tensile response of the non-wovens are greatly inuenced by the ber size distribution in the mats, porosity, individual ber orientation in the mat, berber interaction and entanglement of the bers [20]. These parameters cannot be easily isolated and controlled in the non-woven fabrics. Hence, there has been a remarkable growth and interest in characterizing the tensile deformation behavior of single bers and aligned ber bundles [58]. More recently, it was demonstrated that the size effect is critical in inuencing the ber properties and an abrupt increase in tensile properties is observed at a given average ber diameter [58]. The size effect in the bers is attributed to the process of electrospinning that results in the formation of unique intrinsic structures within the ber geometry [58]. Hence, the focus of this study is to review recent articles that characterize the intrinsic structural properties of the electrospun bers and present possible explanations for the enhanced tensile behavior of the nanobers. Recent articles on electrospinning focused on various spinnable polymeric materials, processing techniques for fabricating nanober assemblies, effects of processing parameters on ber diameter and morphology, characteristics of the bers and their applications [9,12,2123]. However, the inuences of electrospinning on the structure formation and on the tensile strength of the bers are still lacking. To realize the full potential of the bers it is essential to understand the microstructure formation during electrospinning, since the intrinsic structures of the bers affect their overall mechanical deformation behavior. For example, the ordered arrangement of the polymer chains within the ber geometry during electrospinning leads to strengthening of the bers [58]. Thus, the main objective of this review is to outline possible mechanisms that lead to the fabrication of stronger bers and thereby facilitate

Fig. 1. Schematic of the general laboratory setup used for an electrospinning experiment. The inset shows the SEM morphology of the electrospun nylon 6,6 bers. The schematic illustrates the inverted conical path the jet travels before being collected as randomly oriented bers as shown in the inset SEM micrograph. L represents the length of pipette containing the polymer solution and H is the distance between the tip and collector.

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

705

(A)
Polymer solution

(B)

(C)

Taylor cone Pendant drop Induced charges from electric field Jet initiation
Fig. 2. Schematic illustration of the Taylor cone formation: (A) Surface charges are induced in the polymer solution due to the electric eld. (B) Elongation of the pendant drop. (C) Deformation of the pendant drop to the form the Taylor cone due to the charge-charge repulsion. A ne jet initiates from the cone.

an in-depth understanding of the electrospinning process and its role on the microstructure formation. Properties such as molecular orientation and crystallinity of the nanobers and the factors that inuence their deformation behavior are thoroughly analysed. The effects of ber size on the tensile strength and elastic modulus of the bers are also discussed. The review is organized as follows: Sections 2 and 3 discuss the basic concept behind electrospinning and control of ber diameter. Common techniques used to collect controlled morphology of the bers are described in Section 4. The mechanisms which yield controlled morphology, structures and spatial arrangement of the nanobers are critically reviewed. The effects of electrospinning, ber diameter and type of collector used to control the structures, such as crystallinity and molecular orientation, are discussed in Section 5. Various factors that determine the mechanical deformation behavior of the bers are presented in Section 6. Here, the orderly arrangement of ber structures as a function of ber size is emphasized. Finally, potential applications of these bers are given in Section 7. 2. Electrospinning theory Electrospinning or electrostatic spinning is a simple technique which utilizes high electrostatic forces for ber production. Electrospinning, rst introduced by Formhals [24] and later revived by Reneker [3,4], uses high voltage (about 1020 kV) to electrically charge the polymer solution for producing ultra-ne bers (diameters ranging from a few nanometer to larger than 5 lm) [3]. Fig. 1 shows a schematic illustration of the basic electrospinning setup, which essentially consists of a pipette or a syringe lled with polymer solution, a high voltage source and a grounded conductive collector screen. In addition, a metering syringe pump can be used to control the ow rate of the polymer solution. The needle of the syringe typically serves as an electrode to electrically charge the polymer solution and the counter-electrode is connected to the conductive collector screen. Under the inuence of a strong electrostatic eld, charges are induced in the solution and the charged polymer is accelerated towards the grounded metal collector. At low electrostatic eld strength, the pendant drop emerging from the tip of the pipette is prevented from dripping due to the surface tension of the solution [2527]. As the intensity of the electric eld is increased, the induced charges on the liquid surface repel each other and create shear stresses. These repulsive forces act in a direction opposite to the surface tension [28], which results in the extension of the pendant drop into a conical shape and serves as an initiating sur-

Fig. 3. (A) SEM micrograph of the bers showing typical circular morphology and (B) SEM micrograph of the at ribbon structure.

face [2931]. A schematic of the process is shown in Fig. 2. When the critical voltage is reached, the equilibrium of the forces is disturbed and a charged jet emanates from the tip of the conical drop. The discharged jet diameter decreases in size with concomitant increase in length before being deposited on the collector. This process can be explained by the three types of physical instabilities experienced by the jet [25,26]. These instabilities inuence the size and geometry of the deposited bers. The rst instability, also known as the Rayleigh instability is axisymmetric and occurs when the strength of electric eld is low or when the viscosity of the solution is below the optimum value. Use of very low viscosity solutions causes jet break-up and leads to the bead-on-ber morphology. It is attributed to the poor chain entanglement density in the solution and insufcient resistance to the electrostatic eld [31,32]. Rayleigh instability is suppressed at high electric elds (high charge densities) or when using higher concentration of polymer in the solution. Following the initial straight path of the jet, which is controlled by the Rayleigh instability, the polymer jet is inuenced by two other instabilities: the bending and whipping instabilities. These instabilities arise owing to the charge-charge repulsion between the excess charges present in the jet which encourages the thinning and elongation of the jet [25,26]. At high electric forces, the jet is dominated by bending (axisymmetric) and whipping instabil-

706

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

ity (non-axisymmetric), causing the jet to travel in an inverse cone manner. It produces wave or dumb-bell shaped patterns in the jet as shown in Fig. 1. At higher electric elds and at sufcient charge density in the jet, the axisymmetric (i.e., Rayleigh and bending) instabilities are suppressed and the non-axisymmetric instability is enhanced. The whipping instability produces a bending force on the jet, resulting in a high degree of elongation of the jet [32]. During these processes, the solvent evaporates and nally leads to the deposition of ultra-ne bers on the conductive ground electrode. 3. Control of ber diameter Systematic investigations on the effect of electrospinning parameters on the diameter and the morphology of the bers have been reported by several researchers [3335]. Major factors that control the diameter of the bers are: (1) concentration of polymer in the solution, (2) type of solvent used, (3) conductivity of the solution, and (4) feeding rate of the solution. An example of the effect of parameters on ber geometry is shown in Fig. 3. Fig. 3a shows typical circular bers and Fig. 3b shows at ber belts that are obtained because of the rapid evaporation of the solvent. The attened bers are obtained when a fraction of the solvent is trapped inside the ber. When the solvent evaporates, the ber collapses, resulting in at ber belts. Clearly, there is a critical need to produce bers of uniform diameters so that the electrospinning process can be rendered reproducible for scientic modeling and industrial applications. Fridrikh et al. determined the parameters that control the ber diameter using an analytical model [33] that is based on the differences between surface tension of the solution and the electrostatic charge repulsion in the jet. At high electrical eld, the motion of the jet is inuenced by three main forces, namely: (a) external electric eld, (b) normal stresses, which comprise the surface tension and tension resisting the bending of electric eld lines in the jet, and (c) surface charge repulsion. Bending and stretching is a direct effect of normal stresses, which originate from the bending of the centerline of the jet. Hence, the normal stress gives rise to the whipping instability. When the surface charge repulsion exceeds the surface tension, it leads to the whipping instability and bending of the jet. At this stage, the current is constant and consists of conduction and advection current. At the later stage of whipping, the bulk current is dominated by the advection current and the surface charge repulsion is balanced by the surface tension. At this stage, the stretching of the jet is ceased and a constant diameter of the jet is obtained. The developed model predicts the diameter of this terminal jet, assuming that no further thinning of the jet occurs. Thus, the nal diameter of the ber (D) is determined to be a function of surface tension, electric current and surface charge repulsion. The equation for the diameter is:

Fig. 4. Schematic of the rotating drum used for ber collection. The inset SEM micrograph shows the aligned bers obtained using the rotating drum.

cn

Q2 I2

p 2 ln d 3

2 l

!1 3 1

where c is surface tension of the solution, n dielectric constant, Q ow rate of the solution, I current carried by the jet, l initial jet length and d diameter of the nozzle. Primarily, ow rate, electric current and surface tension of the solution control the whipping jet diameter. For instance, increasing the current carrying ability of the jet by 32 times or reducing the ow rate by 32 times results in a ten-fold ber diameter reduction [33,36]. The ow rate of the solution to the nozzle can be easily controlled by using a ow meter. This model is certainly not comprehensive, considering the number of parameters that would control ber diameters. The model, however, neglects the elastic effect due to solvent evapora-

tion and considers the solution Newtonian. The model also neglects the volatility of the solvents and charge carrying ability of the polymers. The accuracy of predicting the diameter of the ber depends on the charge carrying ability of the jet. When non-conductive polymers such as PCL are used for electrospinning, the charges are solely accommodated by the volatile solvent. The charges from the evaporated solvent may reach the collector, which contributes to the measured current [33], and which leads to over-predicting the stretching of the jet. Hence, the model cannot predict the ber diameters accurately for the polymers in a highly volatile solvent. However, theoretical ber diameters of conductive polymers agree well with experimental values. This is due to the fact that the charges stay with the jet until it reaches the collector and drying occurs after the stretching of the jet. Nonetheless, the model provides a simple analytical method to estimate the diameter of the bers with convincing agreement. Eq. (1) evaluates the terminal diameter considering that the collector is stationary. However, further thinning of the bers can be obtained when rotating collectors such as a rotating drum or a rotating disk collector is used [8]. Kotaki et al. [37] showed that the speed of the rotating collector induced tensile stresses on the bers before being wound around the collector. The tensile stresses are responsible for further thinning of the ber diameter, which is not predicted by Eq. (1).

4. Alignment of bers and ber collection methods Recently, it was determined that the nature of the collector inuences signicantly the morphological and the physical characteristics of the spun bers [38,39]. The density of the bers per unit area on the collector and ber arrangement are affected by the degree of charge dissipation upon ber deposition. The most com-

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

707

Fig. 5. Schematic of the disk collector used for ber collection. The SEM micrograph shows the alignment of the bers obtained using the disk collector. Better alignment of the bers is observed compared to the rotating drum.

monly used target is the conductive metal plate that results in collection of randomly oriented bers in non-woven form as shown in Fig. 3a. Liu et al. [38] electrospun cellulose acetate bers on copper mesh, aluminum foil, paper and water as collectors. They found the type of collector used greatly determined the arrangement and packing density of the bers. The use of metal and conductive collectors helped dissipate the charges and also reduced the repulsion between the bers. Therefore, the bers collected are smooth and densely packed. Conversely, the bers collected on the non-conductive collectors do not dissipate the charges which repel each other. Hence, the bers are loosely packed. The bers can also be collected on specially designed collector systems so as to obtain aligned bers or arrays of bers. Recently, researchers focused on achieving highly ordered aligned bers by using mechanical and electrostatic methods to control the electrospinning process. Aligned bers have found importance in many

engineering applications, such as tissue engineering, sensors, nanocomposites, lters, electronic devices [4043]. Some commonly used techniques to align the bers are discussed in the subsections below. 4.1. Rotating drum collector The schematic of the electrospinning setup with a rotating drum collector is shown in Fig. 4. This method is commonly used to collect aligned arrays of bers. Furthermore, the diameter of the ber can be controlled and tailored based on the rotational speed of the drum [4042]. The cylindrical drum is capable of rotating at high speeds (a few 1000 rpm) and of orienting the bers circumferentially. Ideally, the linear rate of the rotating drum should match the evaporation rate of the solvent, such that the bers are deposited and taken up on the surface of the drum. The

Fig. 6. Schematic of static electrodes used for collecting aligned ber bundles. The optical micrograph shows the aligned bers collector using this technique [5].

708

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

alignment of the bers is induced by the rotating drum and the degree of alignment improves with the rotational speed [40,43]. At rotational speeds slower than the ber take-up speed, randomly oriented bers are obtained on the drum. At higher speeds, a centrifugal force is developed near the vicinity of the circumference of the rotating drum, which elongates the bers before being collected on the drum [23,43,44]. However, at much higher speeds, the take-up velocity breaks the depositing ber jet and continuous bers are not collected. 4.2. Rotating disk collector The rotating disk collector is a variation setup of the rotating drum collector and is used to obtain uniaxially aligned bers. Fig. 5 shows the common setup. The advantage of using a rotating disk collector over a drum collector is that most of the bers are deposited on the sharp-edged disk and are collected as aligned patterned nanobers [23,4346]. The jet travels in a conical and inverse conical path with the use of the rotating disk collector as opposed to a conical path obtained when using a drum collector. During the rst stage, the jet follows the usual envelope cone path which is due to the instabilities inuencing the jet. At a point above the disk, the diameter of the loop decreases as the conical shape of the jet starts to shrink. This results in the inverted cone appearance, with the apex of the cone resting on the disk. The electric eld applied is concentrated on the tapered edge of the disk and hence the charged polymer jet is pulled towards the edge of the wheel, which explains the inverted conical shape of the jet at the disk edge. The bers that are attracted to the edge of the disk are wound around the perimeter of the disk owing to the tangential force acting on the bers produced from the rotation of the disk. This force further stretches the bers and reduces their diameter. The quality of ber alignment obtained using the disk is much better than the rotating drum; however, only a small quantity of aligned bers can be obtained since there is only a small area at the tip of the disk. 4.3. Static parallel electrodes The advantage of using this technique lies in the simplicity of the setup and the ease of collecting single bers for mechanical testing. Good alignment has been obtained with this technique. The air gap between the electrodes creates residual electrostatic repulsion between the spun bers, which helps the alignment of the bers [5,4749]. Two non-conductive strips of materials are placed along a straight line and an aluminum foil is placed on each of the strips and connected to the ground as shown in Fig. 6. This technique enables bers to be deposited at the end of the strips so that the bers adhere to the strips in an alternate fashion and

Fig. 8. Structural morphology of electrospun bers displaying the densely packed lamellae and brillar structure. Reprinted with permission from [51]. Copyright [2010], American Institute of Physics.

Fig. 9. Transmission electron micrograph of nylon 6,6 ber displaying preferred orientation of the polymer chains.

collected as aligned arrays of bers. A similar technique by Teo and Ramakrishna [47,49] used double-edge steel blades along a

Fig. 7. Schematic representation of the nanobril in a single POM nanober. The schematic representation of the crystal orientation of 700 nm POM ber is shown and illustrates the conformation of the helical structure of the chain.

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

709

line to collect aligned arrays of bers. The bers were deposited at the gap between the electrodes, however, few bers were found to deposit on the blades. It was resolved by applying a negative voltage between the blades, resulting in the deposition of bers between the blades. 5. Structural properties of electrospun bers Typically, in most semi-crystalline polymers, the bers produced by electrospinning display structural hierarchy. During the ber formation process, a fraction of the chains crystallizes to form lamellae consisting of small crystals and the remaining fraction forms the amorphous phase [5052]. In the presence of shear and elongation forces, the lamellae are organized to form brils and the tie chain molecules pass through the neighboring crystallites to form small-sized bundles. The general structure in the ber is expected as shown in Fig. 7. Due to the shear forces experienced by the jet during electrospinning, the chain orientation (see Fig. 7) aligns along the ber axis [50]. Konkhlang et al. [50] examined the crystal morphology and molecular orientation of polyoxymethylene (POM) ber and found that each ber consists of nanobrils which are aligned parallel to the ber axis. The brils consist of 14 polymer chains and 40 monomeric units. Similar observations have been found by Lim et al. [51]. They visualize the structural morphology of the ber and found the bers to have densely packed lamellae and brillar structures as shown in Fig. 8. The lamellar structures determine the crystallinity of the bers. In between the stacks of lamellae are the relaxed amorphous tie molecules. 5.1. Molecular orientation The polymer jet under the inuence of an electrostatic eld experiences a high degree of elongation strain (104 times the draw ratio and over 106 s1 draw rate). The high elongation strains and shear forces are capable of aligning the macromolecular chains along the ber axis resulting in a high degree of molecular orientation in the bers [58,52]. Zong et al. [53] found the molecular chains in the electrospun PLLA bers to be highly oriented compared to the random-coil shape chains in the PLLA lm. Other polymer bers such as Kevlar display similar orientated chain structures and is observed between the amorphous and the crystalline regions of the bers. Fig. 9 is an example using transmission electron microscopy (TEM) to analyse the chain orientation. The bers are stained with ink so that upon examination with TEM,

there is a phase contrast between the amorphous and crystalline lamellae. Some degree of chain orientation can be found on the surface regions of the bers. Hence, it can be summarized that the process of electrospinning alters the intrinsic structural properties of the material. The orientation extent can be quantied by using X-ray diffraction analysis on the samples. Alternatively, the draw ratio can be used to obtain an estimate of the molecular orientation. It quanties the amount of extension the jet experiences during the electrospinning process [5]. High draw ratios experienced by the jet are capable of aligning the macromolecular chains along the ber axis, thereby inuencing the formation and structure of the crystallites. The draw ratio for spun bers can be calculated as the ratio of the spinning velocity of the collected ber to ejection velocity of the polymer solution from the pipette [5,43,5456]. According to the principle of mass conservation, the velocity of the bers at the ground collector is given by:

mspin

wf 100 pf pr f 2 t

! 2

where mspin is spinning velocity (m/min) when bers are collected at the ground electrode, wf weight (g) of polymer bers on the ground electrode, pf density (g/cc) of the PCL bers, rf average radius (cm) of the collected bers and t electrospinning time (min). Usually, the electrospinning process is run for longer duration of time ($45 min). The ejection velocity of the PCL solution from the pipette is determined from:

msol

wsol 100 psol pr p 2 t

! 3

where msol and wsol are ejecting velocity (m/min) and weight (g) of the polymer solution at a given time during electrospinning, psol density (g/cc) of PCL solution, rp diameter (cm) of the pipette used and t electrospinning time (in minutes). Using Eqs. (2) and (3), the draw ratio is the ratio of mspin/msol [5,43,5456]. The elongation rate of the PCL bers during the electrospinning process is determined from the following equation:

mspin msol
H

where e is elongation rate and H is distance between the pipette and ground collector. Higher draw ratio values are expected to provide better chain orientation in the bers.

0.5
Spun Sample

0.0

Heat Flow (W/g)

Intensity

-0.5

-1.0
Spun Nylon 6,6 Nylon 6,6 pellets (non-spun)

-1.5

10

20

30

40

50

60

-2.0

2
Fig. 10. Typical intensity versus 2h plot obtained from the WAXD analysis on PCL bers. Crystallinity of the bers can be determined as the ratio of the area of the peaks to the total area of the curve [5].

50

100

150

200

250

300

Temperature (C)
Fig. 11. DSC curves of electrospun nylon 6,6 and un-processed (non-spun) nylon 6,6 pellet comparing the melting temperature and heat of fusion.

710

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

5.2. Crystallinity The rapid evaporation of the solvent from the jet accompanied by the rapid structure formation, which occurs within milliseconds ($50 ms) leads to less developed structures in the bers. The rapid solvent evaporation reduces the jet temperature. Thus, the molecules that are aligned along the ber axis have less time to realign themselves, leading to less favorable packing. For most semi-crystalline polymers, the stretched chains under high elongation rate do not get enough time to form crystalline lamellae, which yields lower crystallinity. Hence, the crystallinity in the bers is thereby inuenced by the rate of solvent evaporation [23,57]. The most common technique used to determine the degree of crystallinity is wide angle X-ray diffraction (WAXD) analysis. The ratio of the area under the peaks to the total area under the curve of intensity versus 2h plot is shown in Fig. 10 and it gives the sample crystallinity. Contrary to the theory that electrospinning reduces the crystallinity of the bers, Lee et al. [58] and Reneker et al. [59] reported that the crystalline structure in bers is developed in many polyesters and ductile materials. Moreover, the crystallinity can be even higher than the un-processed polymer pellets. They argue that electrospinning inhibits the development of crystallinity specically for rigid polymers with high glass transition (Tg) values. However, for ductile polymers and polyesters with lower Tg values, such as PCL (Tg $ 60 C), this takes longer time to crystallize. Therefore, ductile polymers have the possibility of crystallization during the jet drawing/elongation process, even after the bers are solidied. Fig. 11 shows the DSC analysis of nylon 6,6 bers compared with the un-processed nylon 6,6 pellets. The results are consistent with the results of Lee et al. [58] and Reneker et al. [59]. The melting enthalpy of electrospun nylon 6,6 is calculated as 107 J/g compared to 91 J/g for the unspun sample, suggesting an increase in the degree of crystallinity.

100

100

80

80

60

60

40

40

20

Crystallinity Molecular Orientation

20

200

400

600

800

0 1000

Fiber Diameter ( m)
Fig. 13. Plots of crystallinity (%) and molecular orientation (%) versus ber diameter for aligned bers. The degree of crystallinity and molecular orientation increases gradually as the ber diameter is reduced [5].

5.3. Effect of ber diameter on structural properties Zussman et al. [60] in their study demonstrate that the electrospun bers possess skin-core morphology. The skin region of the bers contains oriented layered planes that are parallel to the ber axis and contain few crystallites. But the crystallites are misoriented with respect to the ber axis. The properties of the skin differ from those of the core region for the bers as the skin layers are essentially characterized by the oriented layered planes whereas, the core region is characterized by random-coil chains. These results are substantiated by the molecular dynamic simulations of Curgul [61] who has demonstrated that the molecules are oriented preferentially parallel to the surface for the nanobers. The mobility of these chains at the skin is much higher than the mobility of

the chains present in the core region [62,63]; hence the chains at the skin are easily oriented under the inuence of an electric eld. As the ber diameter is reduced, at some critical ber diameter, the size of the skin region becomes comparable to the overall diameter of the ber [6]. Moreover, the oriented layered planes on either side of the ber wall are coupled together and inuence the overall properties of the bers. In contrast, when the ber diameter is increased, orientation of the chains at the surface of the ber walls becomes less compared to the majority of the chains in the ber core region. These results are in agreement with those reported later by Arinstein et al. [6] who have shown that the bers consist of supramolecular region which consists of oriented amorphous chains. As the size of the ber is reduced; the size of the supramolecular structure containing amorphous oriented macromolecules is more signicant compared to the ber diameter. Thus, at the critical ber diameter, the properties of the ber are controlled by the oriented amorphous macromolecules in the supramolecular region. This work [6] is very important and more detailed investigations are still lacking. In a study conducted in our laboratory, we have established the relationship between the microstructure of the PCL bers and their diameter [5]. The degree of crystallinity and molecular orientation in the bers is determined using wide angle X-ray diffraction (WAXD) analysis. Fig. 12 shows the WAXD pattern of the bers with diameters 250 and 900 nm, respectively. The arc width of the strongest equatorial reection provides an indication of the degree of orientation within the samples. It is clear from the WAXD patterns (Fig. 12) that 900 nm-wide bers have lesser degree of orientation compared to 250 nm-wide bers, that is, the wider the bers the less molecular orientation is exhibited. Fig. 13 shows the degree of crystallinity (%) and molecular orientation (%) versus ber diameter. Molecular orientation determined from WAXD increases with decreasing ber diameter. Therefore, it conrms that as the ber diameter is reduced, the alignment of the molecules in the direction of ber axis is improved.

5.4. Effect of collector on the structural properties The type of collector and the speed of the drum/disk collector selected inuence the isotropic or anisotropic alignment of the bers in the mats. Also, the collector type used controls the crystal morphology and molecular orientation [50]. In the article by Kongkhlang et al. [50], they show that when a rotational collector is used, the polymer chains in the crystalline regions are drawn fur-

Fig. 12. WAXD pattern of aligned bers performed on two sets of ber diameter0073: (i) 250 nm and (ii) 900 nm [5].

Molecular Orientation (%)

Crystallinity (%)

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

711

ther in the draw direction compared to the polymer chains in the non-woven fabrics that are obtained using a stationary collector. The force due to the rotational speed of the collector along with the shear and elongation forces may contribute to the alignment of the polymer chains in the direction of the ber axis. Thus, it is expected that the crystal orientation in the bers improves with the speed of the collector [64]. The use of high speed rotational collectors leads to a fanning effect and the evaporation of the solvent is much quicker compared to the stationary collectors [37]. The speed creates a high-viscosity environment for the polymer chains in the electrospinning jet and leads to the transfer of the tensile stress onto the polymer chains during the ber deposition. Thus, the crystallization in the bers occurs due to the sliding diffusion which facilitates formation of extended chain crystals (ECC) from the folded chain crystals by lamellar thickening [37]. It leads to increases in crystal size and crystallinity. This is caused by a more perfect planar zigzag conformation of the ECC structures under the inuence of an applied tensile stress. Also, as the rotational force contributes towards the stretching of the polymer jet, higher rotational speed decreases the diameter of the bers. This explains the ordering of the crystals at higher collector speeds. Furthermore, when the static parallel electrodes are used to obtain aligned ber arrays, the extended chain crystals are not observed from WAXD and infrared spectroscopic analyses. Also, the crystal orientation is expected to be inferior compared to the rotational collectors. Hence, we did not nd a signicant increase in the degree of crystallinity using the parallel electrodes method even though the ber diameter was reduced [5]. 6. Mechanical properties of the bers Polymer nanobers are treated as 1-dimensional systems and have found to possess unusual mechanical properties. The mechanical deformation behavior displayed by the bers is unique and can be signicantly different from their macroscopic counterparts [5,43,37,57,65]. The unique features of the bers are attributed to the process of electrospinning. 6.1. Effect of structural morphology on tensile properties The lamellar and amorphous fractions of the chains within the bers inuence the strength and elastic modulus of the bers. Changes in the structural formation taking place in the bers dur-

ing electrospinning, specically crystallinity and molecular orientation, impart physical uniqueness to the material and play an important role in the deformation behavior of the bers [58]. Hence, knowledge of their intrinsic structures is essential to understand their effects on mechanical properties. The amorphous phase of the bers provides the elastomeric properties and the crystalline phase imparts dimensional stability to the array of molecules [5]. Thus, the mechanical deformation characteristics of the ber is inuenced by the random or/and ordered arrangements of the crystalline and amorphous phases in the ber [58,50,51]. According to Curgul et al. [61], the mechanical deformation of the bers is affected by the skin and core morphologies of the ber. The mass density in the core region is similar to the bulk polymer density. Thus, the core region of the ber exhibits bulk-like structure and physical properties, whereas, the property exhibited by the surface region is entirely different. This is attributed to the signicantly lower density and increased mobility of the chains at the surface/skin of the ber compared to the core region [62,63]. Hence, the overall deformation of the ber is determined by the number of oriented fragments present in the surface regions. This theory is also conrmed by Arienstein et al. [6]. In their study, they concluded that the orientation of the amorphous chains in the supramolecular region of the bers inuences the deformation process of the bers. If this understanding is applied to study the effect of ber diameter on tensile strength, it should result in an exponential increase, or an abrupt shift, in properties as the ber diameter is reduced. This is because the effect of ber surface/skin on the overall nanober is increasingly dominant as (a) the ber surface dimension approaches the radius of gyration of polymer chains, thus constraining the segmental motion, and (b) the ber core region diminishes when the ber diameter decreases. In our previous study [5], we also evaluated the effect of ber diameter on the tensile deformation. The tensile response of the ber was compared with the tensile properties of the bulk polymer system prepared using injection molding. Representative stress strain curves of spun and bulk systems are shown in Fig. 14. There is a signicant difference in the tensile strength and tensile behavior. The stressstrain curve of the spun sample is consistently found not to display the necking phenomenon whereas the bulk sample shows clear necking. This is attributed to the oriented and stretched polymer chains in the spun bers [66,67]. Similar results were obtained by Lu et al. [66].

60

50

Stress (MPa)

40
2

30
Necking

20

10

0 0 2 4 6

Strain (mm/mm)
Fig. 14. Stressstrain curves obtained from tensile tests performed on electrospun PCL and non-spun PCL samples. Curve 1 represents the electrospun sample and Curve 2 represents the non-spun sample [5].

Fig. 15. SEM micrograph of randomly oriented bers with berber fusion points.

712

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

6.2. Effect of collector type on tensile properties 6.2.1. Stationary collector The morphology of the mats obtained using a stationary collector is shown in Fig. 3a. Their mechanical deformation depends greatly on the degree of alignment of the bers within the mat, ber lay-ups and interface properties of berber contact [19,45,66]. Typically, the tensile strength and modulus of the non-woven fabrics are lower than the mats with uniaxially oriented bers. This is attributed to the highly porous nature of the non-woven fabrics. Moreover, during tensile loading, only the bers oriented along the loading direction experience the stretching force, while the bers that are oriented perpendicular to the loading direction do not experience any force. The bers tend to orient in the direction of loading before the non-linear region in the stressstrain curves. After the non-linear point, the ber mesh structure is damaged and better orientation of the bers along the loading direction is observed. This stage is followed by the high orientation of the bers at the maximum stress point and ber breakage at several points is noticed. Such deformation behavior in non-woven fabrics is usually observed when there is no ber-to-ber bonding. The lack of bonding between the bers facilitates easy orientation and stretching of the bers when loaded and can give high degree of elongation before failure [19]. The use of highly volatile solvents during electrospinning can produce non-wovens with little or no ber fusions. Fibers cannot fuse together when the solvent evaporation is high and this also results in weak intermolecular interaction. However, when there is fusion between bers as shown in Fig. 15, the modulus of the non-wovens increases and the elongation to break decreases. The fusion of bers is obtained if the solvent is not completely evaporated during the ber forming process. 6.2.2. Rotational collector Macroscopically aligned bers obtained by modifying the ber collecting system are found to have anisotropic properties [68,69] which can be potentially useful in a variety of optical, mechanical and bio-medical applications. Uniaxial, aligned bers are found to possess anisotropic tensile properties. The tensile strength and modulus of these samples are higher than randomly oriented bers [69]. When the bers are oriented in the loading

direction, the uniaxial orientation of the bers helps the tensile force distribute equally to all bers. Further, since the molecular chains in the bers are aligned along the ber axis, which is in the loading direction, the samples display enhanced strength and modulus. The tensile strength of the aligned ber array samples also depends on the technique used to collect these arrays. When a rotating disk or a drum is used to collect the bers, mechanical forces may be applied to the jet due to the rotational speed of the collector. This force along with the shear and elongation forces entice the alignment and stretching of the polymer chains in the ber axis direction. In addition, the rotational speed of the collector determines the stacking density of the bers. At higher rotational speeds, the deposited bers have a denser lateral packing and minimum inter-ber spacing. Also, the bers tend to have uniform morphology and diameter at higher rotational speeds, which contribute towards the strength of the samples [3946]. 6.3. Effect of ber diameter on tensile properties Fiber structure, geometrical arrangement of the bers, individual ber properties and interaction between bers greatly inuence the mechanical properties of ber mats. These features are difcult to control during the electrospinning process. Therefore, determining the tensile deformation of the single ber is of fundamental importance. Recently, researchers determined the tensile deformation of single bers and demonstrated that the size of the ber inuenced their tensile response. An enhanced behavior is observed by researchers when the diameter of the bers is reduced below the critical diameter [58]. Arinstein et al. [6] in their study demonstrate that the size of the ber has an effect on its deformation behavior. At some critical diameter, the bers display almost an exponential increase in tensile strength. This phenomenon prevails when the size of the supramolecular structures of the bers is comparable with the overall ber diameter. The orientation of macro molecules present in the supramolecular structures of the amorphous phase plays a dominant role to increase the ber mechanical properties. Upon increasing the ber size, both tensile strength and tensile modulus decreases and the larger diameter bers tend to display bulk-like properties. This observation is extremely important for conceivable applications of eletrospun nanobers. Instead of considering such polymers as bers, they can be used as miniaturized high aspect ratio components for devices and sensors. Hence, by acknowledging the abrupt changes in strength and modulus as ber diameter decreases, we cannot use measurements obtained, or extrapolated, from bulk specimens to model devices at nanometer length scale. Fig. 16 shows the tensile strength and tensile modulus versus ber diameter seen in individual PCL bers. The bers with diameters greater than 2 lm do not affect the modulus or tensile strength and can be thought to have bulk-like properties. The enhanced properties of ner diameter bers are attributed to the gradual ordering of the molecular chains and modest increase in the crystallinity of the bers. The size effect can also be due to the densely packed lamellae and brillar structures. In ner diameter bers, the lamellae and brillar structures align themselves along the ber axis, which plays a critical role in enhancing the mechanical properties of the bers. The brillar structure has a high degree of molecular orientation and provides high resistance to the axial tensile force. When the ber diameter is increased, the lamellae tend to re-orientate and the presence of alignment and brillar lamellae structure is decreased, resulting in reduced mechanical properties. Dzenis [67] and others [70,71] modeled this size effect in polymer nanobers and considered the surface energy of the bers to contribute towards the axial tensile force. The assumption made in these studies is that bers of different sizes have similar mor-

400 60

Tensile Strength (MPa)

300 40

200 20
Strength Modulus

0 0.0

0.5

1.0

1.5

2.0
m)

2.5

100 0 3.0

Fiber Diameter (

Fig. 16. Plot of tensile modulus and tensile strength versus ber diameter. Tensile modulus increases with decreasing ber diameter. These results can be attributed to the better molecular orientation and crystallinity in smaller ber diameters. Larger than 2 lm, both tensile modulus and tensile strength appear invariant with ber diameter [5].

Tensile Modulus (MPa)

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

713

Fig. 17. Schematic illustrating the hierarchical organization of bone. It also shows the self-assemble process of mineralization observed in natural composites.

phology and can be considered isotropic. However, the size effect seen in experiments attributes this effect to the electrospinning process and the macromolecular orientations in the bers. The use of an electric eld during electrospinning spontaneously aligns highly mobile molecular chains in the direction of the ber axis, resulting in a higher degree of molecular orientation. The large shearing imposed on the electrospinning jet and the high draw ratio seen for ner bers suggest better chain orientation for the thinner bers. Some studies attribute this effect to the orientation of the chains in the outermost region of the bers [6,7]. Because of the large shear stress and surface tension inuencing the electrospinning jet, it results in a higher number of monomers aligned on the ber surface and in contact with the surface. Besides, as the ber diameter is reduced, the chains orient along both surfaces of the ber and are considered to be physically coupled which enhances the properties. Hence, when the thickness of these surface layers is comparable to the overall ber size, it plays an important role in inuencing the mechanical properties of the bers. In contrast, as the ber size is increased, misorientation of the polymer chains along the surface occurs and the surface layer is no longer comparable to the overall ber size. The higher degree of misorientation present in the ber core region yields bulk-like properties [60,72,73] and dictates the overall ber properties. 7. Prospective applications of electrospun bers 7.1. Fiber composites for tissue engineering applications Ultra-ne bers of biodegradable polymers produced by electrospinning have found potential applications in tissue engineering due to their high surface area to volume ratios and high porosity of the bers [9,14,65,73,74]. Moreover, the exibility of seeding stem cells and human cells on the bers makes electrospun materials most suited for tissue engineering applications [75,76]. The bers produced can be used systematically to design the structures such that they do not only mimic the properties of the extracellular ma-

Fig. 18. SEM micrograph of fracture surface showing the presence of electrospun nanobers in the matrix resin. Reprinted from [16], with permission from Elsevier.

trix (ECM), but also possess high strength and high toughness. For instance, non-woven fabrics exhibit isotropic properties and support neo-tissue formation. These mats resemble the ECM matrix and can be used as skin-scaffold and wound dressing materials where the materials are required to be more elastic than stiff [14,7782]. When anisotropic properties are desired for load-bearing applications, such as musculoskeletal tissues (tendons and ligaments), aligned electrospun bers can be used to mimic the structural anisotropy of the tissues. Many natural polymers (collagen, starch, chitin and chitosan) and synthetic biodegradable polymers (poly(e-caprolactone) (PCL), polylactide (PLA), poly(D, Llactide-coglycolide) (PLGA)) have been widely investigated for potential applications in developing tissue scaffolds [77,8386]. This suggests a thorough understanding of the mechanical behavior of electrospun nanobers is essential. For example, the fracture toughness of synthetic electrospun scaffolds has not been addressed at all and this is a critical factor for assessing the mechanical integrity of the scaffolds. The natural tissues due to

714

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

their hierarchical structural arrangement possess superior fracture toughness values compared to any of the synthetic scaffold materials that are currently used [83]. To match the mechanical integrity of natural composites, the design of the scaffolds developed should mimic the architectural design used by natural tissues. Fig. 17 shows the self-assembly design mechanism seen in bones. Natural tissues at the structural level essentially consists of collagen brils and bone ceramic in the form of hydroxyapatite (HAP). Mineralization of the tissue takes places by the mechanism known as protein ber-guided mineralization. Collagen bers (300 nm in length and 1.23 nm in width) are self-assembled in an orderly fashion and generate channels or grooves. Mineral particles originate and develop at the grooves grow in length and width as sheets of mineral platelets. The mineral platelets are placed parallel to each other and provide strength to the composite. Thus, the arrangement of collagen and bone crystals at the structural level can critically affect the mechanical integrity of the whole system [8890]. The strength of the tissue scaffolds processed using the conventional techniques lack the architecture design seen in the natural composites and hence their mechanical characteristics are drastically different from natural bone composites. However, the electrospinning technique is capable of mimicking the protein guided mechanism and can potentially align the HAP particles in the ber direction. Further, the strength of the bers can be controlled by the loading of HAP llers. Shields et al. [65] showed that electrospun collagen bers with diameters $100 nm intended for articular cartilage repair have modulus $170 MPa and maximum stress of 3.3 MPa. These values closely match the cartilage mechanical properties of Youngs modulus of 130 MPa and maximum stress of $20 MPa. Stanishevsky et al. [87] fabricated composites of hydroxyapatite (HAP)/collagen using electrospinning for hard tissue scaffold applications and demonstrated that the properties of the electrospun material can be easily controlled by the HAP loading in the bers. These results conrm that electrospinning of natural or synthetic polymers for tissue engineering applications are very promising. 7.2. Electrospun ber reinforced composites Although electrospun ber reinforced polymer composites have signicant potential for development of high strength/high toughness materials and materials with good thermal and electrical con-

ductivity, very few studies have investigated the use of electrospun bers in composites [1618,91]. Fig. 18 shows a SEM micrograph of an electrospun ber reinforced composite. Traditional reinforcements in polymer matrices can create stress concentration sites due to their irregular shapes and cracks propagate by cutting through the llers or travelling up, down and around the particles. However, electrospun bers have several advantages over traditional llers [17]. The reinforcing effect of the bers is inuenced essentially by the ber size. Smaller size bers give more efcient reinforcement. Also, as discussed in the previous sections, bers with ner diameters have preferential orientation of the polymer chains along the ber axis. The orientation of macromolecules in the bers improves with the reduction in diameter, making ner diameter bers very strong. Hence, the use of nanometer-sized bers can signicantly enhance the mechanical integrity of the polymer matrix compared to micron-sized bers. Moreover, the high percentage of porosity and irregular pores between the bers can lead to an interpenetrated structure when dispersed in the matrix, which also enhances the mechanical strength due to the interlocking mechanism. These characteristic features of nanobers enable the transfer of applied stress to the bermatrix interface in a better fashion than most of the commonly used ller materials [16]. Current issues related to the use of electrospun nanobers as reinforcement materials are the control of dispersion and orientation of the bers in the polymer matrix. To achieve better reinforcement, electrospun nanobers may need to be collected as a highly aligned yarn instead of a randomly distributed felt so that the post-electrospinning stretching process could be applied to further improve the mechanical properties. Further, if crack growth is transverse to the ber orientation, the fracture toughness of the composite can be optimized. Hence, the interfacial adhesion between bers and matrix material needs to be controlled such that the bers are capable of deecting the cracks by bermatrix interface debonding and ber pull-out. The interfacial adhesion should not be too strong or too weak. Optimal control can only be attained by careful selective ber surface treatment. The dispersion of electrospun mats in the matrix can be improved by trimming the bers to shorter fragments. This can be achieved, if the electrospun bers are collected as aligned bundles (instead of non-woven network), which can then be optically or mechanically trimmed to obtain ber fragments of several 100 nm in length. 7.3. Conductive ber composites Electrospinning has found applications in developing exible and compliant conductive nanobers for applications in miniaturized devices [23,9294]. Researchers seek to develop compliant electrodes for electroactive polymer actuators. Use of electrospinning to produce bers from conductive polymer matrices can be useful for these applications. Moreover, electrospinning can be used to disperse carbon nanotubes (CNT) in bers to improve the mechanical, electrical and conductive properties of the matrix material [10,11]. Due to the high elongation of the polymer jet during electrospinning, the CNTs tend to orient along the ber axis and are embedded in the ber core as shown in Fig. 19. Application of CNTs and carbonaceous llers in polymers is known to improve the electrostatic charge dissipation and electromagnetic shielding efciency, thus improving the overall conductivity of the polymers. Accordingly, many polymers are being investigated that can be easily electrospun and used as matrix material for CNTs. Another advantage of using electrospun bers for developing electrodes is the surface area to volume ratio of the bers. Since the rate of electrochemical reactions is affected by the surface area of the electrode, the high surface area of the bers for the development of porous electrodes can be exploited.

Fig. 19. TEM micrograph of nylon 6,6-CNT ber. The CNTs are embedded in the core region of the ber and are aligned along the ber axis.

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

715

Norris et al. [92] fabricated ultrane electrically conductive polyaniline/polyethylene oxide (PAN/PEO) ber blend using electrospinning. The standard four-point probe method was used to determine the conductivity of non-woven bers and cast lms. By controlling the PAN/PEO ratio in the blend, they improved the conductivity of non-woven mats comparable to that of the cast lm. Ko [95] determined that the size of the ber obtained from conductive polymers has important effects on system response time to electronic stimuli and the current carrying capability of the ber. Using poly(3,4-ethylenedioxythiophene) (PEDOT) brous mats it was demonstrated that the conductivity of the mats increased exponentially as the ber diameter was decreased. Packing density of the molecules in ner diameter bers could be a possible reason for the enhanced conductivity in the bers. It might also be attributed to the intrinsic ber conductivity effect or the geometric surface effect resulting from the reduction in ber diameters. 7.4. Filtration
Fig. 20. TEM micrograph of electrospun ber lled with magnetite particles.

Electrospun bers with micron-sized diameters have found extensive functions in ltration applications [23,9698]. Electrospun non-woven fabrics used for ltration provide the advantages of high surface area to volume ratio, low air resistance, lower lter mass and exibility of adding surface functionality on the bers by blending or incorporating nanollers [97]. Electrospun bers are being widely investigated for aerosol ltration, air cleaning applications in industry and for particle collection in clean rooms. Typically, aerosol particles are ltered due to the electrostatic attraction that exists between the lter media and aerosol particles. Electrospun bers used in ltration media can improve the efciency of ltration as the static charge used to produce the electrospun bers may remain in the bers and help in the ltration of aerosol particles. It is seen that the ltration efciency of the electrospun mats is comparable to the commercially available lters but the advantage lies in the lter mass which is substantially lower for the former than the latter [23,98]. It is well-known that as the surface area of the bers is increased, the surface adhesion properties of the bers improve. Hence, by decreasing the diameter of the ber in the lter media, the efciency of capturing sub-micron sized particles can be significantly improved compared to the larger bers. For efcient ltration, the sizes of the structural elements in the lter media have to match the size of the particles of droplets that are to be captured by the lter media. The advantage of using electrospun bers in the ltration media is that the ber diameters can be easily controlled and can make an impact in high efciency particulate air ltrations. 7.5. Filler reinforced ber systems

larly useful for miniaturized electronic components and load-bearing applications. The key factors that inuence the reinforcing effects of the ller in electrospun bers are the dispersion, distribution and alignment of the llers in the ber matrix. Electrospinning offers an efcient route for obtaining homogeneous dispersion and distribution of the nanoscale reinforcements [10,11]. Moreover, it is demonstrated that the high electrostatic forces and shear force experienced by the jet during electrospinning align the llers along the ber axis. Good dispersion of the llers is essential for constraining the segmental motion of the molecular chains. Further, the use of llers such as CNTs in electrospun bers is seen to align itself along the ber axis (see Fig. 19). The embedded CNT reduces the overall mobility of the polymer chains and provides the connement effect to the neighboring molecules. Thus, orientation of polymer chains during electrospinning and the presence of hard llers within the bers strengthen the bers. In our previous work [5], the reinforcing effect of the HAP ller on a PCL matrix is veried by comparing the tensile strengths of the electrospun ber composites with those of the melt-processed composites. Electrospinning is found to be far superior to meltprocessing. More interestingly, for electrospun ber samples, ller addition increases the tensile strength. However, ller addition decreases the tensile strength of melt-processed composites (bulk). In the electrospun bers, HAP particles are contained in the nanof-

30
Magnetic Fiber

20

Nanoscale reinforcements have been often used by researchers to fabricate multi-functional high strength composites. Therefore, novel brous composites can be obtained by incorporating high strength and high aspect ratio llers in ber matrices [1,10,11,87,99]. For instance, electrospinning is studied for fabricating lightweight brous composite with unique properties for protective clothing and body armor applications [1]. Commonly used llers are carbon nanotubes (CNTs), organoclay, hydroxyapatite (HAP), silica and titania particles. Filler reinforced bers have many potential applications: ultra-strong wires, nanocomposites, nanoprobes, electronic devices, tissue replacement materials [10,11,87] etc. For example, addition of HAP particles to biodegradable polymer bers shows potential for bio-medical applications. Similarly, CNT inclusion in electrospun bers enhances the overall strength and conductivity. Such CNT reinforced bers are particu-

M (emu/g)

10 0 -10 -20 -30 -10000 -5000 0 5000 10000

H (Oe)
Fig. 21. Magnetic hysteresis loop of magnetic ller reinforced electrospun ber.

716

A. Baji et al. / Composites Science and Technology 70 (2010) 703718

ibers and serve to constrain the segmental motion of the polymer chains. Hence, the ber-guided composites are seen to have enhanced tensile strength. It can be concluded that the ber-guided architecture creates a more effective reinforcement compared to the ller-dispersion approach. Thus, the overall mechanical performance of ller reinforced electrospun bers is inuenced by the dispersion and orientation of the llers within the bers. This observation can also be attributed to the re-ordering of molecular chains in electrospun systems which is not seen in polymers processed by melt ow [5].

8. Concluding remarks and future work Since its discovery, signicant advancement has been made in R&D of electrospinning. Researchers have mostly focused on optimizing the processing parameters to obtain bers of desired shapes and forms with little understanding of the parameters that control the enhanced microstructures and mechanical properties. Many polymers have been successfully electrospun into nanobers. However, very few studies can be found on the macromolecular orientation and crystalline structures of the bers. Through this review, concepts behind the structures and morphology of electrospun bers are discussed, their effects on the mechanical properties are emphasized, and future work identied. The unique mechanical properties of the electrospun bers described in this study demonstrate the potential of using these bers for miniaturized polymer devices and composites applications. Recently, researchers identied the size effect of bers on the structural and tensile properties of the bers. For example, tensile properties such as elastic modulus and strength increase with decreasing ber diameter. The enhanced orderliness of the amorphous phase in the supramolecular structure of the ber plays an important role in inuencing the properties of the bers. High shear forces are seen to produce a skin and core morphology in the bers. The heterogeneity in the skin and core regions is established due to the higher degree of chain orientation in the former compared to the latter. The core region of the ber shows bulk-like properties and the skin region displays enhanced properties. Hence, when the skin thickness is comparable to the overall ber diameter, both the tensile modulus and tensile strength are significantly increased. The heterogeneity in the skin and core regions of the bers is more remarkable when the bers are reinforced with CNTs. Inclusion of CNTs in the ber matrix presents an additional interface for surface chain orientation [99]. Therefore, the overall chain orientation increases with the CNT loading. These oriented regions are stiffer compared to the regions of disoriented chains. Consequently, CNT reinforcement leads to stiffening and strengthening of the bers. To-date, there are no extensive experimental studies designed to investigate the molecular structures of the core and skin regions of electrospun bers. The surface behavior at the skin can differ very much from the core properties of the ber. Further investigations on the strengthening mechanisms across the skin and core regions are pressingly needed. Current efforts have been mainly about the incorporation of ller materials to increase the strength of electrospun bers [44,87,99]. However, future work should focus more on the multi-functionality of nanober composites with llers for specic applications. For instance, electrospinning technique can be used to incorporate magnetic llers within the ber matrix to obtain super-paramagnetic nanostructured composite with controlled geometry. Moreover, bers with homogenous dispersion and distribution of the llers are very attractive since the composite is expected to display enhanced magnetic-eld dependent superparamagnetism. Such features are seldom achieved using

other conventional techniques. We have recently demonstrated that magnetic particle electrospun bers can display superparamagnetism at ambient temperature. A TEM micrograph of one such composite with uniformly distributed magnetic particles is shown in Fig. 20 [100] and its magnetic hysteresis loop at 300 K is plotted in Fig. 21 [100]. In addition, the composite bers are also seen to deect in the direction of increasing magnetic eld and conrm eld responsive behavior. Further, the incorporation of magnetic particles enhances the ber elastic modulus. Thus, many attractive features like mechanical strength, magnetic and conductive properties of the nanoparticles can be utilized to obtain multi-functional composites. Developments of such nanostructured ber composites can be especially useful in miniaturized electronic parts and for electromagnetic interference shielding. Also, the use of biodegradable polymeric ber as the carrier matrix can enhance its usefulness in bio-medical, magnetic resonance imaging and drug-delivery applications. Acknowledgments We would like to thank the Australian Research Council for the support of this work. SCW acknowledges the nancial support from the US National Science Foundation under the CAREER Award CMMI #0746703 and Award DMI #0520967. Sincere thanks are due to TA Blackledge and DH Reneker for helpful discussions and constructive comments during the preparation of this review paper. References
[1] Ko FK, Sukigara S, Gandhi M, Ayutsede J. Electrospun carbon nanotube reinforced silk bers. US patent no. 0082197; 2007. [2] Zarkoob S, Eby RK, Reneker DH, Hudson SD, Ertley D, Adams WW. Structure and morphology of electrospun silk nanobers. Polymer 2004;11:39737. [3] Reneker DH, Chun I. Nanometre diameter bres of polymer, produced by electrospinning. Nanotechnology 1996;7:21623. [4] Doshi J, Reneker DH. Electrospinning process and applications of electrospun bers. J Electrostat 1995;35:15160. [5] Wong SC, Baji A, Leng SW. Effect of ber diameter on tensile properties of electrospun poly(-caprolactone). Polymer 2008;21:471322. [6] Arinstein A, Burman M, Gendelman O, Zussman E. Effect of supramolecular structure on polymer nanober elasticity. Nat Nanotech 2007;2:5962. [7] Tan EPS, Ng SY, Lim CT. Tensile testing of single ultrane polymeric ber. Biomaterials 2005;26:14536. [8] Zussman E, Burman M, Yarin AL, Khaln R, Cohen Y. Tensile deformation of electrospun nylon-6, 6 nanobers. J Polym Sci Part B: Polym Phys 2006;44:14829. [9] Lannutti J, Reneker D, Ma T, Tomasko D, Farson D. Electrospinning for tissue engineering scaffolds. Mater Sci Eng (Biomim Supramol Sys) 2007;27:5049. [10] Dror Y, Salalha W, Khaln RL, Cohen Y, Yarin AL, Zussman E. Carbon nanotubes embedded in oriented polymer nanobers by electrospinning. Langmuir 2003;19:701220. [11] Salalha W, Dror Y, Khaln RL, Cohen Y, Yarin AL. Zussman E Single-walled carbon nanotubes embedded in oriented polymeric nanobers by electrospinning. Langmuir 2004;20:98525. [12] Deitzel JM, Kleinmeyer J, Harris D, Tan NCB. The effect of processing variables on the morphology of electrospun nanobers and textiles. Polymer 2001;42:26172. [13] Gibson P, Schreuder-Gibson H, Rivin D. Transport properties of porous membranes based on electrospun nanobers. Colloid Surf A-Physicochem Eng Asp 2001;187:46981. [14] Li WJ, Laurencin CT, Caterson EJ, Tuan RS, Ko FK. Electrospun nanobrous structure: a novel scaffold for tissue engineering. J Biomed Mater Res 2002;60:61321. [15] Matthews JA, Wnek GE, Simpson DG, Bowlin GL. Electrospinning of collagen nanobers. Biomacromolecules 2002;3:2328. [16] Fong H. Electrospun nylon 6 nanober reinforced BIS-GMA/TEGDMA dental restorative composite resins. Polymer 2004;45:242732. [17] Tian M, Gao Y, Liu Y. Bis-GMA/TEGDMA dental composites reinforced with electrospun nylon 6 nanocomposite nanobers containing highly aligned brillar silicate single crystals. Polymer 2007;48:27208. [18] Kim JS, Reneker DH. Mechanical properties of composites using ultrane electrospun bers. Polym Compos 1999;20:12431. [19] Zhang M, Atkinson KR, Baughman RH. Multifunctional carbon nanotube yarns by downsizing an ancient technology. Science 2004;306:135861. [20] Wei X, Xia Z, Wong SC, Baji A. Modelling of mechanical properties of electrospun nanobre network. Int J Exp Comp Biomech 2009;1:4557.

A. Baji et al. / Composites Science and Technology 70 (2010) 703718 [21] Agarwal S, Greiner A, Wendorff JH. Electrospinning of manmade and biopolymer nanobers progress in techniques, materials, and applications. Adv Funct Mater 2009;19:286379. [22] Dzenis Y. Material science: spinning continuous bers for nanotechnology. Science 2004;304:19179. [23] Huang ZM, Zhang YZ, Kotaki M, Ramakrishna S. A review on polymer nanobers by electrospinning and their applications in nanocomposites. Comp Sci Technol 2003;63:222353. [24] Formhals, A. Apparatus for producing articial laments from material such as cellulose acetate. US patent no. 1975504; 1934. [25] Reneker DH, Yarin AL, Fong H, Sureeporn K. Bending instability of electrically charged liquid jets of polymer solutions in electrospinning. J Appl Phys 2000;87:453147. [26] Shin YM, Hohman MM, Brenner MP, Rutledge GC. Experimental characterization of electrospinning: the electrically forced jet and instabilities. Polymer 2001;42:995567. [27] Shin YM, Hohman MM, Brenner MP, Rutledge GC. Electrospinning: a whipping uid jet generates submicron polymer bers. Appl Phys Lett 2001;78:114951. [28] Fang X, Reneker DH. DNA bers by electrospinning. J Macromol Sci Phys 1997;B36:16973. [29] Taylor GI. Disintegration of water drops in an electric eld. Proc Roy Soc Lond Ser A 1969;313:453. [30] Taylor GI. Electrically driven jets. Proc Roy Soc Lond Ser A 1966;291:145. [31] Hohman MM, Shin M, Rutledge G. Electrospinning and electrically forced jets. I. Stability theory. Phys Fluids 2001;13:220120. [32] Zuo WW, Zhu MF, Yang W, Yu H, Chen YM, Zhang Y. Experimental study on relationship between jet instability and formation of beaded bers during electrospinning. Polym Eng Sci 2005;45:7049. [33] Fridrikh SV, Yu JH, Brenner MP, Rutledge GC. Controlling the ber diameter during electrospinning. Phys Rev Lett 2003;90:14. [34] Tao J, Shivkumar S. Molecular weight dependent structural regimes during the electrospinning of PVA. Mater Lett 2007;61:23258. [35] Eda G, Shivkumar S. Bead structure variations during electrospinning of polystyrene. J Mater Sci 2006;41:57048. [36] Ko FK. Nanober technology in nanotubes and nanobers (advanced materials). Yury Gogotsi, editor. Taylor and Francis Group LLC 2006. [37] Yee WA, Nguyen AG, Lee PS, Kotaki M, Liu Y, Tan BT. Stress-induced structural changes in electrospun polyvinylidene diuoride nanobers collected using a modied rotating disk. Polymer 2008;49:4196203. [38] Liu H, Hsieh YL. Ultrane brous cellulose membranes from electrospinning of cellulose acetate. J Polym Sci Part B: Polym Phys 2002;40:211929. [39] Kim HS, Kim K, Jin HJ, Chin IJ. Morphological characterization of electrospun nano-brous membranes of biodegradable poly(L-lactide) and poly(lactideco-glycolide). Macromol Symp 2005;224:14554. [40] Sundaray B, Subramanian V, Natarajan TS, Xiang RZ, Chang CC, Fann WS. Electrospinning of continuous aligned polymer bers. Appl Phys Lett 2004;84:12224. [41] Katta P, Alessandro M, Ramsier RD, Chase GG. Continuous electrospinning of aligned polymer nanobers onto a wire drum collector. Nano Lett 2004;4:22158. [42] Theron A, Zussman E, Yarin AL. Electrostatic eld-assisted alignment of electrospun nanobres. Nanotechnology 2001;12:38490. [43] Fennessey SF, Farris RJ. Fabrication of aligned and molecularly oriented electrospun polyacrylonitrile nanobers and the mechanical behavior of their twisted yarns. Polymer 2008;45:421725. [44] Mathew G, Hong JP, Rhee JM, Lee HS, Nah C. Preparation and characterization of properties of electrospun poly(butylene terephthalate) nanobers lled with carbon nanotubes. Polym Test 2005;24:7127. [45] Xu CY, Inai R, Kotaki M, Ramakrishna S. Aligned biodegradable nanobrous structure: a potential scaffold for blood vessel engineering. Biomaterials 2004;25:87786. [46] Dersch R, Liu T, Schaper AK. Electrospun nanobers: internal structure and intrinsic orientation. J Polym Sci: Part A: Polym Chem 2003;41:54553. [47] Teo WE, Ramakrishna S. Electrospun bre bundle made of aligned nanobres over two xed points. Nanotechnology 2005;16:187884. [48] Kim GH, Kim WD. Formation of oriented nanobers using electrospinning. Appl Phys Lett 2006;88:233101. [49] Teo WE, Ramakrishna S. A review on electrospinning design and nanobre assemblies. Nanotechnology 2006;17:89106. [50] Konkhlang T, Tashiro K, Kotaki M, Chirachanchai S. Electrospinning as a new technique to control the crystal morphology and molecular orientation of polyoxymethylene nanobers. J Am Chem Soc 2008;130:154606. [51] Lim CT, Tan EPS, Ng SY. Effects of crystalline morphology on the tensile properties of electrospun polymer nanobers. Appl Phys Lett 2008;92:141908. [52] Baji A, Wong SC, Blackledge TA, Leng S. Mechanical properties of electrospun composites and their crystallinity measurements using X-ray diffraction. ANTEC Conf Proc 2008;12:1424. [53] Zong XH, Kim K, Fang DF, Ran SF, Hsiao BS, Chu B. Structure and process relationship of electrospun bioabsorbable nanober membranes. Polymer 2002;43:440312. [54] Brazinsky I, Williams AG, LaNieve HL. The dry spinning process: comparison of theory with experiment. Polym Eng Sci 1975;15:83441.

717

[55] Wang H, Shao H, Hu X. Structure of silk broin bers made by an electrospinning process from a silk broin aqueous solution. J Appl Polym Sci 2006;101:9618. [56] White JL, Hancock TA. Fundamental analysis of the dynamics, mass transfer, and coagulation in wet spinning of bers. J Appl Polym Sci 1981;21:315770. [57] Greiner A, Wendorff JH. Electrospinning: a fascinating method for the preparation of ultrathin bers. Angew Chem Int Edit 2007;46:5670703. [58] Lee KH, Kim HY, La YM, Lee DR, Sung NH. Inuence of a mixing solvent with tetrahydrofuran and N, N-dimethylformamide on electrospun poly(vinyl chloride) nonwoven mats. J Polym Sci Part B: Polym Phys 2002;40:225968. [59] Reneker DH, Kataphinan W, Theron A, Zussman E, Yarin AL. Nanober garlands of polycaprolactone by electrospinning. Polymer 2002;43:678594. [60] Prilutsky S, Zussman E, Cohen Y. The effect of embedded carbon nanotubes on the morphological evolution during the carbonization of poly(acrylonitrile) nanobers. Nanotechnology 2008;19:165603 (9pp). [61] Curgul S, Vliet KJV, Rutledge GC. Molecular dynamics simulations of size dependent structural and thermal properties of polymer nanobers. Macromolecules 2007;40:84839. [62] Ngai KL. Interpreting the dynamics of nano-conned glass-formers and thin polymer lms: importance of starting from a viable theory for the bulk. J Polym Sci Part B Polym Phys 2006;44:298095. [63] Ngai KL. Mobility in thin polymer lms ranging from local segmental motion, Rouse modes to whole chain motion: a coupling model consideration. Eur Phys J E 2002;8:22535. [64] Kim WH, Lee KH, Khil MS, Ho SY, Kim HY. The effect of molecular weight and the linear velocity of drum surface on the properties of electrospun poly(ethylene terephthalate) nonwovens. Fiber Polym 2004;5:1227. [65] Shields KJ, Beckman MJ, Bowlin GL. Mechanical properties and cellular proliferation of electrospun collagen type II. Tiss Eng 2004;10:15107. [66] Lu JW, Zhang ZP, Ren XZ, Chen YZ, Yu J, Guo ZX. High-elongation ber mats by electrospinning of polyoxymethylene. Macromolecules 2008;41:37624. [67] Wu XF, Dzenis YA. Size effect in polymer nanobers under tension. J Appl Phys 2007;102:044306. [68] Thomas V, Jose MV, Chowshury S, Sullivan JF, Dean DR, Vohra YK. Mechano-morphological studies of aligned nanobrous scaffolds of polycaprolactone fabricated by electrospinning. J Biomater Sci Polym Ed 2006;17:96984. [69] Huang C, Chen S, Reneker DH, Lai CL, Hou HQ. High-strength mats from electrospun poly(p-phenylene biphenyltetracarboximide) nanobers. Adv Mater 2006;18:66871. [70] Cuenot S, Fretigny CH, Demoustier CS, Nysten B. Surface tension effect on the mechanical properties of nanomaterials measured by atomic force microscopy. Phys Rev B 2004;69:165410. [71] Dingreville R, Qu JM, Cherkaoui M. Surface free energy and its effect on the elastic behavior of nano-sized particles, wires and lms. J Mech Phys Solid 2005;53:182754. [72] Chew SY, Hufnagel TC, Lim CT, Leong KM. Mechanical properties of single electrospun drug-encapsulated nanobers. Nanotechnology 2006;17: 388091. [73] Chen F, Peng X, Li T, Chen S, Wu XF, Reneker DH, et al. Mechanical characterization of single high-strength electrospun polyimide nanobres. J Phys D: Appl Phys 2008;41:025308. [74] Chen M, Patra PK, Warner SB, Bhowmick S. Optimization of electrospinning process parameters for tissue engineering scaffolds. Biophys Rev Lett 2006;1: 15378. [75] Wang YZ, Blasioli DJ, Kim HJ, Kim HS, Kaplan DL. Cartilage tissue engineering with silk scaffolds and human articular chondrocytes. Biomaterials 2006;27: 443442. [76] Moroni L, Licht R, de Boer J, de Wijn JR, van Blitterswijk CA. Fiber diameter and texture of electrospun PEOT/PBT scaffolds inuence human mesenchymal stem cell proliferation and morphology, and the release of incorporated compounds. Biomaterials 2006;27:491122. [77] Yoshimoto H, Shin YM, Terai H, Vacanti JP. A biodegradable nanober scaffold by electrospinning and its potential for bone tissue engineering. Biomaterials 2003;24:207782. [78] Powell HM, Supp DM, Boyce ST. Inuence of electrospun collagen on wound contraction of engineered skin substitutes. Biomaterials 2008;29:83443. [79] Smith D, Reneker DH. PCT/US00/27737; 2001. [80] Kenawy ER, Bowlin GL, Manseld K, Layman J, Simpson DG, Sanders EH, et al. Release of tetracycline hydrochloride from electrospun poly(ethyleneco-vinylacetate), poly(lactic acid), and a blend. J Control Release 2002;81: 5764. [81] Kim K, Luu YK, Chang C, Fang DF, Hsiao BS, Chu B, et al. Incorporation and controlled release of a hydrophilic antibiotic using poly(lactide-coglycolide)-based electrospun nanobrous scaffolds. J Control Release 2004;98:4756. [82] Goldberg M, Langer R, Jia XQ. Nanostructured materials for applications in drug delivery and tissue engineering. J Biomat Sci Polym Ed 2007;18: 24168. [83] Baji A, Wong SC, Srivatsan TS, Njus GO, Matur G. Processing methodologies for polycaprolactone-hydroxyapatite composites: a review. Mater Manuf Process 2006;21:2118. [84] Min BM, Jeong L, Nam YS, Kim JM, Kim JY, Park WH. Formation of silk broin matrices with different texture and its cellular response to normal human keratinocytes. Int J Biol Macromol 2004;34:22330.

718

A. Baji et al. / Composites Science and Technology 70 (2010) 703718 [93] Chroanakis IS, Grapenson S, Jakob A. Conductive polypyrrole nanobers via electrospinning: electrical and morphological properties. Polymer 2006;47:1597603. [94] MacDiarmid AG, Jones WE, Norris ID, Gao J, Johnson AT, Pinto NJ, et al. Electrostatically-generated nanobers of electronic polymers. Synthetic Met. 2001;19:2730. [95] Ko FK. Nanober technology: bridging the gap between nano and macro world in nanoengineered nanobrous materials. Netherlands: Kluwer Academic Publishers; 2004. [96] Yoon K, Hsiao BS, Chu B. Functional nanobers for environmental applications. J Mater Chem 2008;18:532634. [97] Heikkila P, Taipale A, Lehtimaki M, Harlin A. Electrospinning of polyamides with different chain compositions for ltration application. Polym Eng Sci 2008;48:116876. [98] Barhate RS, Loon CK, Ramakrishna S. Preparation and characterization of nanobroud ltering media. J Membr Sci 2006;283:20918. [99] Ji Y, Li C, Wang G, Koo J, Ge S, Li B, et al. Connement-induced super strong PS/ MWNT composite nanober. EPL 2008;84 [article no. 56002]. [100] Baji A, Mai Y-W, Wong SC. Unpublished work.

[85] Boudriot U, Goetz B, Dersch R, Greiner A, Wendorff JH. Role of electrospun nanobers in stem cell technologies and tissue engineering. Macromol Symp 2005;225:916. [86] Boland ED, Telemeco TA, Simpson DG, Wnek GE, Bowlin GL. Utilizing acid pretreatment and electrospinning to improve biocompatibility of poly(glycolic acid) for tissue engineering. J Biomed Mater Res Part B 2004;71:14452. [87] Stanishevsky A, Chowdhury S, Chinoda P, Thomas V. Hydroxyapatite nanoparticle loaded collagen ber composites: microarchitecture and nanoindentation study. J Biomed Mater Res 2008;86:87382. [88] Curry JD. Proc Roy Soc Lond Ser B 1977;196:44363. [89] Stupp SI, Brawn PV. Molecular manipulation of microstructures: biomaterials, ceramics, and semiconductors. Science 1997;277:12428. [90] Stupp SI, LeBonheur V, Walker K, Li LS, Huggins KE, Keser M. Supramolecular materials: self-organized nanostructures. Science 1997;276:3849. [91] Bergshoef MM, Vancso GJ. Transparent nanocomposites with ultrathin, electrospun nylon-4, 6 ber reinforcement. Adv Mater 1999;11:13625. [92] Norris ID, Shaker MM, Ko FK, MacDiarmid AG. Electrostatic fabrication of ultrane conducting bers: polyaniline/polyethylene oxide blends. Synthetic Met 2000;114:10914.

Das könnte Ihnen auch gefallen