A review in electrospinning

Embed Size (px)

Citation preview

  • 8/10/2019 A review in electrospinning

    1/8

    Formation of fibers by electrospinning

    Gregory C. Rutledge , Sergey V. Fridrikh

    Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

    Received 20 December 2006; accepted 14 April 2007

    Available online 22 August 2007

    Abstract

    Electrostatic fiber formation, also known as electrospinning, has emerged in recent years as the popular choice for producing continuousthreads, fiber arrays and nonwoven fabrics with fiber diameters below 1 m for a wide range of materials, from biopolymers to ceramics. It

    benefits from ease of implementation and generality of use. Here, we review some of the basic aspects of the electrospinning process, as it is

    widely practiced in academic laboratories. For purposes of organization, the process is decomposed into five operational components: fluid

    charging, formation of the cone-jet, thinning of the steady jet, onset and growth of jet instabilities that give rise to diameter reduction into the

    submicron regime, and collection of the fibers into useful forms. Dependence of the jetting phenomenon on operating variables is discussed.

    Continuum level models of the jet thinning and jet instability are also summarized and put in some context.

    2007 Elsevier B.V. All rights reserved.

    Keywords: Electrospinning; Nanofiber; Electrohydrodynamics

    Contents

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1384

    2. Brief description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385

    3. Fluid charging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385

    4. Cone-jet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385

    5. Slender thinning jet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1386

    6. Jet instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1388

    7. Fiber collection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390

    8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390

    Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390

    1. Introduction

    Electrospinning is a term used to describe a class of fiber

    forming processes by which electrostatic forces are employed to

    control the production of fibers. It is closely related to the more

    established technology of electrospraying, which generally refers

    to processes in which electrostatic forces are used to control the

    formation of droplets. Spinningin this context is a textile term

    that derives from the early use of spinning wheels to form yarns

    from natural fiber staples like cotton and is commonly used to

    identify fiber-forming processes for synthetic fibers as well. In

    both electrospinning and electrospraying, the role of the

    electrostatic forces is to supplement or replace the conventional

    mechanical forces (e.g. hydrostatic, pneumatic) used to form the

    jet and to reduce the size of the fibers or droplets, hence the term

    electrohydrodynamic jetting. Electrospraying was described in

    the technical literature by Zeleny as early as 1914 [1], while

    electrospinning first appears in the patent literature in 1902 [2].

    Electrospraying has enjoyed a rich history in the intervening

    Available online at www.sciencedirect.com

    Advanced Drug Delivery Reviews 59 (2007) 1384 1391www.elsevier.com/locate/addr

    Corresponding author. Tel.: +1 617 253 0171.

    E-mail address:[email protected](G.C. Rutledge).

    0169-409X/$ - see front matter 2007 Elsevier B.V. All rights reserved.doi:10.1016/j.addr.2007.04.020

    mailto:[email protected]://dx.doi.org/10.1016/j.addr.2007.04.020http://dx.doi.org/10.1016/j.addr.2007.04.020mailto:[email protected]
  • 8/10/2019 A review in electrospinning

    2/8

    century, with commercial applications in ink jet printing, mass

    spectrometry, and fuel injection, to name a few. Electrospinning

    on the other hand has developed more slowly. Notable early

    studies are those by Baumgarten [3], and by Larrondo and St.

    John Manley [4] but apparently neither received subsequent

    follow-up. In the 1990's Reneker and co-workers [5,6], drew

    attention to electrospinning as a means to produce small diameter,continuous filaments. In contrast to conventional synthetic fiber

    forming processes such as those used for high speed spinning of

    nylon or polyester, where continuous fibers ranging from 10 to

    500 m are produced, electrospinning readily leads to the

    formation of seemingly continuous fibers with diameters ranging

    from0.01 to 10m. The literature on electrohydrodynamic jetting

    and atomization for electrosprays is extensive, and relevant to

    electrospinning as well [79]. This brief review is intended to

    provide an overview of the electrospinning process, a subjective

    assessment of the primary operational variables as they are

    understood at the current time, and some of the mathematical

    modeling that has been instrumental in developing our under-standing of the process. Due to its brevity and the rapid growth of

    the field in the past few years, it is necessarily incomplete, as our

    understanding of the technology continues to improve with the

    benefit of on-going research.

    2. Brief description

    In its simplest incarnation, the process of electrospinning is

    relatively easy to implement. In the laboratory, a glass pipette is

    filled with a solution comprising the spin dopeand mounted at

    an angle sufficient to prevent discharge of the fluid from the

    pipette under its own weight. A wire electrode inserted into the

    fluid reservoir in the pipette and charged to a relatively highvoltage, typically around 10 kV, provides the necessary charging

    potential. Under such conditions, the fluid itself charges to a high

    potential (positive or negative, depending on polarity of the

    voltage generator). The typical spin dope has sufficient

    conductivity for the induced charge to relax to a free surface or

    interface on a time scale short compared to the experimental time

    scale, but it otherwise acts as a dielectric. The term leaky

    dielectric is descriptive of such fluids [10]. The repulsion

    between charges at the free surface then works against surface

    tension and fluid elasticity to deform the droplet into a conicalshape, called a Taylor cone after the pioneering work of G.I.

    Taylor[11]. Beyond a critical charge density, this cone is unstable

    and a jet of fluid is emitted from the tip of the cone. This charged

    jet then seeks a path to ground. As it does so, the fluid filament is

    accelerated and stretched and may experience any of a number of

    instabilities, the relative importance of which depends on

    numerous variables, including the properties of the fluid, the

    electric field environment and the dynamical behavior of the

    jetting phenomenon. A suitable collector electrode is used to

    direct this path to ground. Thus, the electrospinning process may

    be broken down into several operational components: (i) charging

    of the fluid, (ii) formation of the cone-jet, (iii) thinning of the jet inthe presence of an electric field, (iv) instability of the jet, and (v)

    collection of the jet (or its solidified fibers) on an appropriate

    target. A simple schematic of the process is shown inFig. 1.

    3. Fluid charging

    In electrospinning, generation of charge within the fluid

    usually occurs by virtue of contact with and flow across an

    electrode held at high (positive or negative) potential, referred to

    as induction charging[12]. Depending on the nature of the fluid

    and the polarity of the applied potential, free electrons, ions or

    ion pairs may be generated as charge carriers in the fluid; the

    generation of charge carriers can be very sensitive to solutionimpurities. The formation of ions or ion pairs by induction

    results in the formation of an electrical double layer. In the

    absence of flow, the double layer thickness is determined by the

    ion mobility in the fluid; in the presence of flow, ions may be

    convected away from the electrode and the double layer

    continually replenished. Charging of the fluid in electrospinning

    is typically field-limited, with the break-down field strength in

    dry air being on the order of 30 kV/cm between flat plates.

    Inductive charging is generally suitable for fluids with

    conductivities of the order of 102 S/m. For nonconductive

    fluids such as hydrocarbons and polymer melts, charge may be

    injected directly into the fluid, as is done in electrostaticthrusters, by the use of two electrodes, one having a needle-like

    geometry. An early example of this was reported by Kim and

    Turnbull for electrospraying [13]. For electrospinning, Kelly

    has reported the production of fibers using a charge injection

    device originally developed for fuel atomization[14]. Relative-

    ly high throughputs were achieved in the latter case, but the

    resulting jet deviates from the conventional cone-jet/thinning jet

    mode described in detail below.

    4. Cone-jet

    Cloupeau and Prunet-Froche [15] have described several

    functioning modes of operation for electrohydrodynamic jets

    Fig. 1. Schematic of the electrospinning process. (a) high voltage power supply;

    (b) charging device; (c) high potential electrode (e.g. flat plate); (d) collector

    electrode (e.g. flat plate); (e) current measurement device; (f) fluid reservoir; (g)flow rate control; (h) cone; (i) thinning jet; (j) instability region.

    1385G.C. Rutledge, S.V. Fridrikh / Advanced Drug Delivery Reviews 59 (2007) 13841391

  • 8/10/2019 A review in electrospinning

    3/8

    leading ultimately to the formation of charged droplets or aerosol.

    For the case of electrospinning, the cone-jet regime is of particular

    note since it represents the initial stage of formation of the fiber.

    Available information on charge injection devices suggests that

    these often form what Cloupeau and Prunet-Foche callramified

    jets. At sufficiently high voltages, multiple jetting from the

    spinneret may also be observed in the conventional electrospin-ning process [8,16], but is not discussed further here. Tracer

    particles have been used to reveal an axisymmetric circulation

    flow within the cone, suggesting that the stable jet forms by

    shearingfluid off the surface of the cone[9]. Ganan-Calvo[17]

    has developed an asymptotic solution for the cone-jet configu-

    ration that provides expressions for the electric current, cone-jet

    shape, and charge distribution within the jet. Yan et al.[18]have

    modeled the steady jet equations of the cone-jet numerically as a

    2-dimensional free surface flow.

    5. Slender thinning jet

    As it emits from the cone, the fluid jet forms a slender,

    continuous liquid filament. The fluid is charged and accelerates

    in the presence of the electric field created by the high potential

    of the spinneret and by the charged fluid itself. Fig. 2. shows

    images of several representative jets for different flow rates ( Q)

    and applied electric fields (E). Numerous attempts have been

    made to decipher trends in electrospun jet operation as functions

    of fluid and operating parameters under experimental control.

    Within certain limits, for a given fluid there exist ranges of

    driving voltage (V) and flow rate over which the electrospinning

    can be maintained stably for long periods. The electrospinning

    process constitutes an electrical circuit, so one can measure as

    well the current (I) flowing through this circuit. Numerousinvestigators have reported currentI increasing with voltage V

    in a nonlinear fashion[16,1923]. Theron et al. and Demir et al.

    fit their data to power laws, IVx, with values ofx =2.17 and

    2.7, respectively, while Deitzel et al. and Shin et al. attributed

    the nonlinear behavior to the onset of fluid instabilities which

    they observed to occur at a critical voltage. Shin et al. [22]and

    Demir et al. [16] also showed that for sufficiently large flow

    rates, currentIis roughly proportional to flow rate Q, indicative

    of a volumetric charge density,I/Q, independent of flow rate. At

    low flow rates,I/Qtends to increase asIbecomes less sensitive

    to Q [24]. Theron et al., on the other hand, reported IQx,

    1.04bb0.26, for a range of fluids, which they attributed to

    ion mobility-limited charging of the fluid in their equipment

    configuration[23].

    Mathematical description of the thinning liquid jet can be

    formulated within the context of conventional electrohydrody-

    namics[17]. One such exposition is that presented by Hohman

    et al.[25]for Newtonian fluids, which makes use of the slender

    body approximation to write perturbative expansions in the

    aspect ratio of the jet for the relevant jet characteristics, such as

    diameter (h), velocity (v), surface charge density () and local

    electric field strength (E), in terms of radial (r) and axial (z)

    coordinates. These are substituted into the equations for

    conservation of mass (for an incompressible fluid), conserva-tion of charge, differential momentum balance and the

    electrostatic field arising from a line of charge whose potential

    is expressed by Coulomb's law. The resulting equations are then

    truncated after the leading order terms, leading to a relatively

    simple 1D model for the slender thinning jet. The relevant fluid

    properties, all of which can be characterized independently of

    the electrohydrodynamic model, are viscosity (), density (),

    surface tension (), conductivity (K) and dielectric constant ().

    Operational parameters for the electrospinning process, within

    this model, are flow rate (Q), electric current (I), and applied

    field (E

    ), the last of which arises as a consequence of the

    applied voltage (V) and inter-electrode distance (D), E

    = V/D.

    The model was shown to describe well the jet shape for lowconductivity fluids. For high conductivity fluids, convergence

    of the equations to a solution was found to be laborious, a

    problem associated with specifying the (unknown) initial

    surface charge density on the fluid jet exiting the spinneret.

    Alternatively, Hartman [26] proposed a different boundary

    Fig. 2. Cone-jets of 2% solutions of polyethylene oxide (MW=920 k g/mol) in water. D =45 cm in all cases. (a)Q = 0.02 ml/min,E=0.282 kV/cm; (b)Q =0.10 ml/min,E=0.344 kV/cm; (c) Q = 0.50 ml/min,E= 0.533 kV/cm; (d) Q = 1.00 ml/min, E=0.716 kV/cm. Images courtesy of Dr. Jian Yu.

    1386 G.C. Rutledge, S.V. Fridrikh / Advanced Drug Delivery Reviews 59 (2007) 13841391

  • 8/10/2019 A review in electrospinning

    4/8

    condition with specified initial slope (dh/dz)0 at the nozzle.

    Feng [27] subsequently reformulated the slender body model

    for Newtonian jets using an approximation for the electric field

    equation that, fortuitously, circumvents the convergence

    problems of Hohman et al. The resulting steady state equations

    can then be written as follows[27]:

    h2v 1 1

    Eh2 Pehvr 1 2

    vv0 1

    Fr

    3

    Re

    1

    h2 gh2v0 0

    1

    We

    h0

    h2

    X rr0 bEE0 2Er

    h

    3

    E El

    ln v rh

    0

    b

    2 Eh

    2 00 4

    where the relevant dimensionless groups (indicative of the

    relative strengths of different terms) are:

    electric Peclet number charge convection=conduction:

    Pe2

    P

    e v0

    Kh05

    Froude number: inertia=gravity Fr v20gh0

    6

    Reynolds numberinertia=viscosity: Re

    qv0h0

    g 7

    Weber number:inertia=surface tension Weqv20h0

    g8

    Aspect ratio: v D

    h09

    Electrostatic force parameter electrostatic=inertia

    XP

    eE20qv20

    10

    Charge induction beP

    e1 11

    and the corresponding characteristic values employed are h0forlength,v0 Q= ph

    20

    for velocity, E0 I= ph

    20K

    for electric

    field and 0=E0for charge density.is the dielectric constant

    of the outer fluid (typically, air in conventional electrospinning).

    Subscript 0 is used to indicate values taken at the nozzle.

    Gravitational and aerodynamic contributions were analyzed by

    Reneker et al. and determined to be inconsequential for the

    steady electrospun jet[28].

    In general, the results of Hohman et al. and of Feng indicate

    that the shape of the jet is strongly dependent upon the evolution

    of surface charge density and local electric field. Concurrent

    with the initial rapid reduction in jet diameter, the latter two

    quantities rise quickly to maximum values as charges relax to

    the jet surface and surface advection current becomes more

    important relative to bulk conduction current. The characteristic

    length over which the initial dramatic thinning of the jet takes

    place can be identified with the axial distance where advectionand conduction currents are equal. Analysis of these equations

    (Fridrikh, unpublished) leads to the following relation for this

    nozzle regime length L:

    L5 K4Q7q3 ln v 2

    8p2El

    I5P

    e 2 12

    Beyond this characteristic length, the jet thins more slowly.

    Sufficiently far from the nozzle (circa 30h0), the jet approaches

    the asymptotic regime where all terms except electrostatic and

    inertial must eventually die out. In this asymptotic regime, the

    following scaling relations for jet diameter, charge density and

    local electric field[29,30]have been derived:

    h Q3q

    2p2El

    I

    14

    z14 13

    r I3q

    32p2El

    Q

    14

    z14 14

    EEl

    IQq

    2El

    12 lnv

    4pP

    e

    z

    32 15

    The scaling expression for jet diameter was first derived by

    Kiricheneko et al.[29]and more generally for pseudoplastic and

    Fig. 3. Schematic illustration of perturbations associated with several of the

    lowest order instabilities, distinguished by their azimuthal wave number, s. Top

    views illustrate cross-sections of the jet at maximum amplitudes of (oscillatory)

    perturbation, with bold and dashed contours representing different positions

    along the jet length. Bottom views illustrate changes in shape and centerline

    down the length of the jet.+/ are used to indicate regions of positive or negative

    deviation from the unperturbed jet shape. Perturbations are exaggerated beyond

    the linear instability regime. (a) unperturbed cylindrical fluid element; (b)

    varicose (s =0) instability; (c) whipping (s =1) instability (also called bending

    or kink instability in the literature); (d) splitting (s =2) instability. Growth of

    the varicose instability leads to equal sized droplets; growth of the splitting

    instability leads to two equal sized sub-jets. Higher order (sN2) instabilities are

    also conceivable.

    1387G.C. Rutledge, S.V. Fridrikh / Advanced Drug Delivery Reviews 59 (2007) 13841391

  • 8/10/2019 A review in electrospinning

    5/8

    dilatant fluids by Spivak and Dzenis [31]. It was confirmed

    experimentally by Shin et al.[22].

    In between the nozzle regime and asymptotic regime,

    viscosity may play a role in determining the rate of thinning

    of the jet. A balance between electrostatic and viscous terms

    (Fridrikh, unpublished) suggests the following scaling in this

    regime:

    h 6gQ2

    pEl

    I

    12

    z1 16

    Feng [27] and more recently Carroll and Joo [32] have

    furthermore extended the slender jet model to include

    viscoelastic effects. Here too, the primary role of extensional

    thickening is to alter the shape of the jet in the intermediate

    regime of the thinning jet, resulting in sharply thinning cone-jets

    such as that shown in Fig. 2a. Recent experimental studies of

    model Boger fluids by Yu et al. confirm the rapid initial thinning

    of the cone-jet, for viscoelastic fluids, and suggest theemergence of a new regime where the scaling of diameter

    with distance along the jet increases from the inverse 1/4 power,

    typical of Newtonian fluids, towards the inverse 1/2 power for

    fluids with strong extensional thickening[33].

    Although illuminating, these analytical solutions for the steady

    jet behavior are limited by the breakdown of the slender jet

    approximation in the nozzle regime, where the cone-jet occurs.

    Numerical methods such as that employed by Yan et al. [18]for

    the cone-jet can be matched to the slender thinning jet solutions.

    6. Jet instability

    In principle, the cone-jet operation is sufficient to draw outcontinuous fibers to very small diameter by electrospinning.

    However, the fluids typically used in electrospinning do not

    always solidify sufficiently en route to the collector to remain

    fibrous after impact on the collector. In practice, as the jet thins,

    it ultimately succumbs to one or more fluid instabilities which

    distort the jet as they grow. A family of such instabilities exists

    and can be analyzed for different conditions of the symmetry

    (axisymmetric or nonaxisymmetric) of the growing perturba-

    tions of the jet, using linear instability analysis [34,35].Fig. 3

    illustrates the perturbations associated with several of the lowest

    order instabilities. Although splaying of the jet was proposed in

    the 1990's as the primary mode of instability [5], and someevidence has been found for splitting in post mortem

    micrographs of fibers[36]and for secondary jetting in images

    of jets[21,28](a particularly beautiful example appears in the

    book by Frankel[37]), such events are relatively rare; the most

    common mode of instability in electrospun jets appears to be the

    growth of lateral excursions of the jet, the so-called whipping

    mode[3,28,38]. This phenomenon is illustrated inFig. 4for jets

    of water and polymer solution; it should be noted that this

    instability is not a consequence of viscoelasticity of the fluid

    and may occur for Newtonian[25,39]as well as non-Newtonian

    fluids. The main competing mode of instability, at least for the

    relatively dilute solutions often employed to achieve the

    smallest fiber diameters in electrospinning, is that of droplet

    break-up, the mode that leads to electrospraying. For the steady

    thinning jets described above, Hohman et al.[25]performed the

    linear instability analysis for both droplet break-up and

    whipping modes and produced operating diagrams for

    electrospinning that illustrate the combinations of controllable

    parameters (flow rate, Q, and applied field, E

    = V/D) under

    which a particular fluid will electrospin (whipping) rather thanelectrospray (droplet breakup). The full dispersion relation for

    the whipping instability of viscous, charged fluids in a finite

    electric field was reported by Hohman et al.[25]. In the vicinity

    of the onset of whipping, charge repulsion is relatively strong

    compared to electrical shear stress, and surface tension effects

    are relatively minor, allowing one to neglect the contributions to

    the dispersion relation arising from the electric field, surface

    tension and finite conductivity. In this approximation, consid-

    erable simplification of the dispersion relation is obtained

    [40,41]:

    xk4 3qQh2

    4p3Reh0 k2

    2pr2

    P

    e 32ln v x2q: 17

    where is the growth rate and k is the wave number of the

    perturbation. Solving for and differentiating this with respect

    tok, one obtains the growth rate and wave number for the most

    unstable (i.e. fastest growing) mode:

    xmax 9p5r4h0Re

    8h2q2QP

    e 2 2ln v 3 2

    1=318

    kmax p3rh0Re

    h2qQ 1=3

    2pqP

    e 2ln v 3

    1=6

    19

    These equations reflect the destabilizing effect of charge

    repulsion (strongest for short wave length perturbations) and the

    stabilizing effect of viscosity (also strongest for short wave

    length perturbations), as well as the damping of the instability

    growth rate due to inertia.

    With respect to droplet breakup, surface stresses due to

    charge-charge repulsion and due to interaction of charges on the

    fluid with the applied electric field serve to destabilize the

    whipping instability relative to droplet breakup, making the

    former the more important mode at high field strength and

    charge density. As a caveat, however, it should be borne in mind

    that multiple instabilities can grow simultaneously, and even awhipping jet can ultimately decay into droplets, as illustrated by

    the water jet inFig. 4. Of particular note, Hohman et al. reported

    a second, electrically driven droplet breakup instability that

    becomes important at high electric fields in fluids of finite

    conductivity, in addition to the conventional, surface tension-

    driven instability. Hohman et al. interpreted their equations for

    both droplet breakup and whipping instabilities in terms of their

    underlying physical mechanisms for a number of limiting cases.

    From the equation of motion reported by Hohman [25],

    Fridrikh et al. [24] also obtained an expression for the lateral

    growth of the jet excursions arising from the whipping

    instability far from its onset, deep in the nonlinear regime.

    This equation contains terms due to acceleration of the charged

    1388 G.C. Rutledge, S.V. Fridrikh / Advanced Drug Delivery Reviews 59 (2007) 13841391

  • 8/10/2019 A review in electrospinning

    6/8

  • 8/10/2019 A review in electrospinning

    7/8

    presence can have benefits for selected applications, such as

    wettability [49]. At least one strategy to overcome this

    phenomenon involves stabilization of the jet against the droplet

    breakup by introducing a more viscoelastic outer fluid in a

    two-fluid co-axial jet configuration[5052]. The resulting core-

    shell fiber morphology may be of interest in its own right;

    alternatively, either the shell or core may be removed to yieldsolid or hollow fibers, respectively. On a related note, Gupta and

    Wilkes [53] have reported a side-by-side configuration for a

    two-fluid spinneret.

    7. Fiber collection

    Electrospun fibers are charged in flight, as described above,

    and travel downfield until impact with a lower potential

    electrode, or collector. A particular strength of the electropin-

    ning process is that fiber collection may be guided by

    manipulation of the electric field lines. The most straightfor-

    ward configuration to analyze is the parallel electrodeconfiguration described by Taylor [7] and subsequently

    modified for electrospinning by Shin [36] and illustrated in

    Fig. 1 earlier. In this configuration, static parallel plate

    electrodes are used at both the nozzle and the collector; this

    configuration ensures parallel field lines throughout the process

    zone, while the field configuration around the nozzle helps to

    suppress corona discharge. A similar observation was made

    using a ring electrode near the nozzle[19]. For the simplest case

    of a static, planar collection electrode, the conventional

    electrospinning process is axially symmetric and provides no

    preference of direction for the fibers upon collection, resulting

    in a nonwoven membrane of unoriented fibers. The integrity of

    the nonwoven fabric depends on the random overlapping offibers, interactions between fibers, and welding that may

    occur at contact points between fibers if they are still somewhat

    fluid when collected. However, the axial symmetry of the

    process is easily overcome by a number of strategies, thus

    resulting in different degrees of fiber orientation in the final yarn

    or membrane product. One such strategy involves use of a

    rotating drum[54]or moving belt collector to produce oriented

    fiber membranes. A rotating wheel with beveled edge has been

    used successfully to collect fibers into an oriented, multifila-

    ment thread[55]. Shaped collectors such as pins [56], parallel

    bars[57], or even simple wire meshes can be used to direct fiber

    collection preferentially, as the charged fibers tend to formbridges between pins or bars when appropriately spaced. In this

    manner, small area arrays of highly oriented fibers have been

    obtained. Rotating one of a pair of parallel disks or rings can be

    used to produce short sections of twisted yarn[58]. For more

    complicated collector shapes, build-up of the polymer mem-

    brane on portions of the collector can alter the field lines to

    ensure conformal coating of the object. Introduction of

    secondary ring electrodes can be used to prolong the steady

    jet regime and to focus the fiber deposition, as mentioned earlier

    [42]. For the rare case where fiber instabilities are completely

    absent, oriented fibers can be collected by writing, for

    example on the face of a rotating wheel[59]. Conductive liquid

    nonsolvents have also been used as the collector, resulting in a

    wet-spinning variant of the electrospinning process and

    production of fibers from solutions of nonvolatile liquids[6,60].

    8. Conclusion

    The main attractions of electrospinning to date have been the

    ease of implementation on a lab scale and the very small diameterfibers that are readily obtained. As emphasized by Dzenis [61],

    electrospinning is a top-down manufacturing process, which

    offers certain advantages to fiber formation in terms of cost and

    productivity over more complicated bottom-up approaches.

    Fibers diameters between 100 nm and 1 m are most typical,

    although fibers with an average diameter as small as 30 nm [62]

    and as large as 10 m have been reported. This is two to three

    orders of magnitude smaller thanconventional, commercial fibers

    (10500 m) and the typical human hair (ca. 100 m).

    Furthermore, in contrast to other forms of nanofibers (e.g. carbon

    nanotubes), electrospun fibers are essentially continuous. This is

    supported by both modeling and observations; fiber ends arerarely observed in micrographs of fibers obtained during steady

    state operation. The continuous nature of the fibers is potentially

    important for occupational health reasons.

    Small diameter implies that a large fraction of the material is

    near the surface of the fiber, and molecules embedded within the

    fiber encounter shorter diffusion path lengths to the fiber

    surface. Commensurate with the reduction in diameter of these

    fibers, the resulting nonwoven membranes typically have very

    high specific surface area, high porosity and small pore size

    (pore size being equated here with some characteristic distance

    between fibers) compared to other fibrous materials. These

    qualities make electrospun materials attractive as candidates for

    tissue engineering scaffolds, drug delivery systems, catalyst andenszyme supports, sensors, and other applications in biology,

    medicine and controlled release.

    Acknowledgements

    The authors are grateful to the National Textile Center

    (Project #'s M98-D01 and MD01-D22) and the US Army

    Institute for Soldier Nanotechnologies at MIT (ARO contract

    DAAD-19-02-D0002) for funding and to our several colla-

    borators and team members for enlightening discussions.

    References

    [1] J. Zeleny, Phys. Rev. 1914, 3, 69; Proc. Camb. Philos. Soc. 1915, 18, 17;

    Phys. Rev. 1917, 10, 1.

    [2] W.J. Morton, US Patent 705,691, 1902.

    [3] P.K. Baumgarten, J. Colloid Interface Sci. 36 (1971) 71.

    [4] L. Larrondo,R. St. John Manley, J. Polym.Sci.: Polym.Phys. 19 (1981)909;

    J. Polym. Sci.: Polym. Phys. 19 (1981) 921; J. Polym. Sci.: Polym. Phys.

    19 (1981) 933.

    [5] J. Doshi, D.H. Reneker, J. Electrost. 35 (1995) 151.

    [6] G. Srinivasan, D.H. Reneker, Polymer Int. 36 (1995) 195.

    [7] G.I. Taylor, Proc. R. Soc. Lond. A313 (1969) 453.

    [8] I. Hayati, A.I. Bailey, T.F. Tadros, J. Colloid Interface Sci. 117 (1987) 205.

    [9] I. Hayati, A.I. Bailey, T.F. Tadros, J. Colloid Interface Sci. 117 (1987) 222.

    [10] G.I. Taylor, Proc. R. Soc. Lond., Ser. A 291 (1966) 159.[11] G.I. Taylor, Proc. R. Soc. A280 (1964) 383.

    1390 G.C. Rutledge, S.V. Fridrikh / Advanced Drug Delivery Reviews 59 (2007) 13841391

  • 8/10/2019 A review in electrospinning

    8/8

    [12] A.G. Bailey, Electrostatic Sprayingof Liquids, Wiley and Sons, New York,

    1988.

    [13] K. Kim, R.J. Turnbull, J. Appl. Phys. 47 (1976) 1964.

    [14] Kelly, A.J., U.S. Patent Application 2001/0046599 A1, Nov. 29, 2001.

    [15] M. Cloupeau, B. Prunet-Foch, J. Aerosol Sci. 25 (1994) 1021.

    [16] M.M. Demir, I. Yilgor, E. Yilgor, B. Erman, Polymer 43 (2002) 3303.

    [17] A.M. Ganan-Calvo, Phys. Rev. Lett. 79 (1997) 217.

    [18] F. Yan, B. Farouk, F. Ko, J. Aerosol Sci. 34 (2003) 99.[19] R. Jaeger, M.M. Bergshoef, C.M.I. Batlle, H. Schnherr, G.J. Vancso,

    Macromol. Symp. 127 (1998) 141.

    [20] H. Fong, D.H. Reneker, Electrospinning and the formation of nanofibers,

    in: D.R. Salem, M.V. Sussman (Eds.), Structure Formation in Polymeric

    Fibers, Hanser, 2000.

    [21] J.M. Deitzel, J.D. Kleinmeyer, D. Harris, N.C. Beck Tan, Polymer 42

    (2001) 261.

    [22] Y.M. Shin, M.M. Hohman, M.P. Brenner, G.C. Rutledge, Polymer 42

    (2001) 9955.

    [23] S.A. Theron, E. Zussman, A.L. Yarin, Polymer 45 (2004) 2017.

    [24] S.V. Fridrikh, J .H. Yu, M.P. Brenner, G.C. Rutledge, Phys. Rev. Lett. 90

    (2003) 144502.

    [25] M.M. Hohman, Y.M. Shin, G.C. Rutledge, M.P. Brenner, Phys. Fluids 13

    (2001) 2201; Phys. Fluids 13 (2001) 2221.

    [26] R.P.A. Hartman, D.J. Brunner, D.M.A. Camelot, J.C.M. Marijnissen, B.Scarlett, J. Aerosol Sci. 30 (1999) 823.

    [27] J.J. Feng, Phys. Fluids 14 (2002) 3912.

    [28] D.H. Reneker, A.L. Yarin, H. Fong, S. Khoombhongse, J. Appl. Phys. 87

    (2000) 4531.

    [29] V.N. Kirichenko, I.V. Petryanov-Sokolov, N.N. Suprun, A.A. Shutov, Sov.

    Phys. Dokl. 31 (1986) 611.

    [30] M.M. Hohman, Ph. D. Thesis, University of Chicago, 2000.

    [31] A.F. Spivak, Y.A. Dzenis, Appl. Phys. Lett. 73 (1998) 3067.

    [32] C.P. Carroll, Y.L. Joo, Phys. Fluids 18 (2006) 053102.

    [33] J.H. Yu, S.V. Fridrikh, G.C. Rutledge, Polymer 47 (2006) 4789.

    [34] D.A. Saville, Phys. Fluids 13 (1970) 2987; Phys. Fluids 14 (1971) 1095;

    J. Fluid Mech. 48 (1971) 815.

    [35] V.Y. Shkadov, A.A. Shutov, Fluid Dyn. Res. 28 (2001) 23.

    [36] M.M. Bergshoef, G.J. Vancso, Adv. Mater. 11 (1999) 1362.

    [37] F. Frankel, Envisioning Science: The Design and Craft of the ScienceImage, MIT Press, Cambridge, MA, 2002, pp. 136137.

    [38] Y.M. Shin, M.M. Hohman, M.P. Brenner, G.C. Rutledge, Appl. Phys. Lett.

    78 (2001) 1149.

    [39] R.H. Magarvey, L.E. Outhouse, J. Fluid Mech. 13 (1962) 151.

    [40] A.L. Yarin, S. Khoombongse, D.H. Reneker, J. Appl. Phys. 89 (2001)

    3018.

    [41] S.V. Fridrikh, J.H. Yu, M.P. Brenner, G.C. Rutledge, in: H Fong, D.H.

    Reneker (Eds.), Polymer Nanofibers, American Chemical Society Sympo-

    sium Series, vol. 918, American Chemical Society, Washington DC, 2006.[42] J.M. Deitzel, J.S. Kleinmeyer, J.K. Hirvonen, N.C. Beck Tan, Polymer 42

    (2001) 8163.

    [43] J.S. Stephens, S. Frisk, S. Megelski, F.J. Rabolt, D.B. Chase, Appl.

    Spectrosc. 55 (2001) 1287.

    [44] R. Jaeger, H. Schnherr, G.J. Vancso, Macromolecules 29 (1996) 7634.

    [45] M. Goldin, J. Yerushalmi, R. Pfeffer, R. Shinnar, J. Fluid Mech. 38 (1969)

    689.

    [46] A.L. Yarin, Free Liquid Jets and Films: Hydrodynamics and Rheology,

    Longman, New York, 1993.

    [47] J. Eggers, Rev. Mod. Phys. 69 (1997) 865.

    [48] H. Fong, I. Chun, D.H. Reneker, Polymer 40 (1999) 4585.

    [49] M. Ma, Y. Mao, M. Gupta, K.K. Gleason, G.C. Rutledge, Macromolecules

    38 (2005) 9742.

    [50] Z. Sun, E. Zussman, A.L. Yarin, J.H. Wendorff, A. Greiner, Adv. Mater. 15

    (2003) 1929.[51] D. Li, Y. Xia, Nano Lett 4 (2004) 933.

    [52] J.H. Yu, S.V. Fridrikh, G.C. Rutledge, Adv. Mater. 16 (2004) 1562.

    [53] P. Gupta, G.L. Wilkes, Polymer 44 (2003) 6353.

    [54] B. Ding, H.Y. Kim,S.C. Lee, D.R. Lee, K.J.Choi, FibersPolym.3 (2002)73.

    [55] A. Theron, E. Zussman, A.L. Yarin, Nanotechnology 12 (2001) 384.

    [56] B. Sundaray, V. Subramanian, T.S. Natarajan, R.Z. Xiang, C.C. Chang, W.S.

    Fann, Appl. Phys. Lett. 84 (2004) 12221224.

    [57] D. Li, Y. Wang, Y. Xia, Nano Lett. 3 (2003)1167; Adv. Mater. 16 (2004)361.

    [58] P.D. Dalton, D. Klee, M. Mller, Polymer 46 (2005) 611.

    [59] J. Kameoka, H.G. Craighead, Appl. Phys. Lett., 3 (2003) 371.

    [60] M.S. Khil, S.R. Bhattarai, H.Y. Kim, S.Z. Kim, K.H. Lee, J. Biomed.

    Mater. Res. B 72 (2005) 117.

    [61] Y. Dzenis, Science 304 (2004) 1917.

    [62] A.G. MacDiarmid, W.E. Jones, I.D. Norris, J. Gao, A.T. Johnson, N.J.

    Pinto, J. Hone, B. Han, F.K. Ko, H. Okuzaki, M. Llaguno, Synth. Met. 119(2001) 27.

    1391G.C. Rutledge, S.V. Fridrikh / Advanced Drug Delivery Reviews 59 (2007) 13841391