electrohydrodynamic quenching in polymer melt electrospinning

9
Electrohydrodynamic quenching in polymer melt electrospinning Eduard Zhmayev, Daehwan Cho, and Yong Lak Joo Citation: Phys. Fluids 23, 073102 (2011); doi: 10.1063/1.3614560 View online: http://dx.doi.org/10.1063/1.3614560 View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v23/i7 Published by the American Institute of Physics. Related Articles Dynamics of magnetic chains in a shear flow under the influence of a uniform magnetic field Phys. Fluids 24, 042001 (2012) Travelling waves in a cylindrical magnetohydrodynamically forced flow Phys. Fluids 24, 044101 (2012) Two-dimensional numerical analysis of electroconvection in a dielectric liquid subjected to strong unipolar injection Phys. Fluids 24, 037102 (2012) Properties of bubbled gases transportation in a bromothymol blue aqueous solution under gradient magnetic fields J. Appl. Phys. 111, 07B326 (2012) Magnetohydrodynamic flow of a binary electrolyte in a concentric annulus Phys. Fluids 24, 037101 (2012) Additional information on Phys. Fluids Journal Homepage: http://pof.aip.org/ Journal Information: http://pof.aip.org/about/about_the_journal Top downloads: http://pof.aip.org/features/most_downloaded Information for Authors: http://pof.aip.org/authors Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Upload: yong

Post on 06-Aug-2016

215 views

Category:

Documents


0 download

TRANSCRIPT

Electrohydrodynamic quenching in polymer melt electrospinningEduard Zhmayev, Daehwan Cho, and Yong Lak Joo Citation: Phys. Fluids 23, 073102 (2011); doi: 10.1063/1.3614560 View online: http://dx.doi.org/10.1063/1.3614560 View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v23/i7 Published by the American Institute of Physics. Related ArticlesDynamics of magnetic chains in a shear flow under the influence of a uniform magnetic field Phys. Fluids 24, 042001 (2012) Travelling waves in a cylindrical magnetohydrodynamically forced flow Phys. Fluids 24, 044101 (2012) Two-dimensional numerical analysis of electroconvection in a dielectric liquid subjected to strong unipolarinjection Phys. Fluids 24, 037102 (2012) Properties of bubbled gases transportation in a bromothymol blue aqueous solution under gradient magneticfields J. Appl. Phys. 111, 07B326 (2012) Magnetohydrodynamic flow of a binary electrolyte in a concentric annulus Phys. Fluids 24, 037101 (2012) Additional information on Phys. FluidsJournal Homepage: http://pof.aip.org/ Journal Information: http://pof.aip.org/about/about_the_journal Top downloads: http://pof.aip.org/features/most_downloaded Information for Authors: http://pof.aip.org/authors

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Electrohydrodynamic quenching in polymer melt electrospinning

Eduard Zhmayev, Daehwan Cho, and Yong Lak Jooa)

School of Chemical and Biomolecular Engineering, Cornell University, Ithaca, New York 14853, USA

(Received 29 September 2010; accepted 29 June 2011; published online 26 July 2011)

Infrared thermal measurements on polymer melt jets in electrospinning have revealed rapid

quenching by ambient air, an order of magnitude faster than predicted by the classical Kase and

Matsuo correlation. This drastic heat transfer enhancement can be linked to electrohydrodynamic

(EHD) effects. Analysis of EHD-driven air flow was performed and included into a comprehensive

model for polymer melt electrospinning. The analysis was validated by excellent agreement of

both predicted jet radius and temperature profiles with experimental results for electrospinning of

Nylon-6 (N6), polypropylene (PP), and polylactic acid (PLA) melts. Based on this analysis, several

methods that can be used to inhibit or enhance the quenching are described. VC 2011 AmericanInstitute of Physics. [doi:10.1063/1.3614560]

I. INTRODUCTION

Electrospinning utilizes applied electric field to form

and drastically accelerate charged fluid jets. It is an effective

technology to produce nanofibers for applications requiring

high surface area to mass ratios. Nanofibers have already

been successfully applied to high efficiency filtration and

will find numerous uses in medical, energy, textile, and

chemical sensing fields.1–3 To date, polymer solution electro-

spinning has been used almost exclusively due to its simplic-

ity, but it has been recently shown that submicron scale

fibers can also be electrospun from polymer melts,4–7 which

is commercially attractive due to environmental and eco-

nomic benefits of the solvent-free process. However, careful

process optimization and control is required to obtain sub-

micron fibers. In melt systems, this is accomplished by

management of extensional viscosity and solidification via

the thermal environment.

Numerous studies on heat transfer in conventional melt

spinning8–12 have served as a basis for thermal management

in melt electrospinning. However, significant differences

between the two processes, especially in the jet thinning rate,

warranted a dedicated study. In this paper, we will discuss

the physics governing heat transfer from the polymer jet to

the quenching ambient air in melt electrospinning, and then

support the arguments by comparison of experimental

thermal and radius profiles for various polymer melts to the

theoretical model predictions.

II. EXPERIMENTAL

The melt electrospinning experimental setup, Figure 1,

and procedure have been described in detail in our previous

publications.7,13–15 A number of temperature control features

have been incorporated, with special attention dedicated to

the melt reservoir and nozzle heaters. Ability to separately

control the spinning region ambient temperature and the

collector has been included, although these features were

intentionally disabled in this study to limit interference with

thermal measurements and to suppress the lateral motion of

the jet. The air-quenched jet temperature profiles were cap-

tured using non-contact thermal imaging with a FLIR Ther-

maCAM EX320 infrared (IR) camera. The jet diameter

profiles were obtained using a Redlake MotionPro HS-3 high

speed camera, as described in literature.13–15 Electrospinning

studies were carried out using 24-gauge needles (0.292 mm

ID) for spinnerets, 29 kV applied potential, and 0.09 m dis-

tance between the spinneret and the collector. The power

supply by Gamma High Voltage Research, Inc., ES30P-

20W, is rated for 0 to 30 kV, with maximum output current

of 660 lA.

Three different polymer melts Nylon-6 (N6), polypro-

pylene (PP), and polylactic acid (PLA) have been studied.

The N6 pellets were provided by Hyosung, Inc., Korea, and

have a melt flow index (MFI) of 3. The 129 kDa isotactic PP

melt was supplied by Clarcor Corporation. The 186 kDa

PLA pellets were purchased from NatureWorks LLC. The

nozzle temperatures and the corresponding flow rates for

various experiments are presented in Table I. The material

properties have been determined in previous studies and are

summarized in Table II. IR thermal measurements are

strongly dependent on the material emissivity, which is a

function of material, temperature, and jet thickness. Obtain-

ing the correct values for emissivity is challenging and is

typically accomplished by equipment calibration at known

conditions. In this work, emissivity for each polymer was

obtained by heating the polymer pellets on a hot plate to a

temperature representative of the electrospinning process

and then calibrating the thermal camera reading by varying

the emissivity. The resulting emissivities and calibration

temperatures are included in Table II.

Due to equipment limitations, neither the coronating

electrode current nor the resulting air velocities could be

measured accurately in experiments. The corona output cur-

rent needed for our model was estimated by extrapolation of

numerical calculations17 for a similar needle-plate geometry

and was found to be �221 lA at 29 kV, Figure 2, which is

within the power supply specifications and is in line with the

experimentally estimated power output. The air velocities

along the polymer jet could not be measured due to strong

a)Author to whom correspondence should be addressed. Electronic mail:

[email protected].

1070-6631/2011/23(7)/073102/8/$30.00 VC 2011 American Institute of Physics23, 073102-1

PHYSICS OF FLUIDS 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

electric field interference. The air velocities away from the

spinline axis decay rapidly, as noted in the numerical

study,17 and in experiments could not be distinguished from

the usual laboratory ambient air circulation. Therefore, no

direct validation for the calculated air velocities could be

presented. Instead, the resulting polymer jet temperature and

radius profiles were used for in-direct validation.

III. THEORY AND MODEL

A. Electrospinning heat transfer model

Electrospinning is a convoluted problem, in which heat

transfer has to be treated in conjunction with momentum and

mass conservation, electric field equations, viscoelastic

effects, and, in some cases, in-flight crystallization. These

equations have been incorporated in our comprehensive melt

electrospinning model, which is utilized as the framework

for present investigations. For the details on derivation and

validation of the electrospinning model, interested readers

are referred to publications.13–15 Here, we will focus on the

heat transfer aspects.

The conservation of energy equation was obtained by

the steady state shell balance on a fluid element in the

Lagrangian frame. Axial conduction was neglected, since it

is dominated by the axial convective heat transport.

Ultimately, the change in internal energy of the fluid element

is balanced by viscous heating, heat released due to crystalli-

zation, and radial convection to the surrounding cooling air:

qCpVdT

dz¼ szz � srrð Þ dV

dz� qDHf V

dhdz� 2h T � T1ð Þ

R; (1)

where q, Cp, and DHf are the polymer melt density, heat

capacity, and heat of fusion; V, T, R, and h are the average

cross-sectional jet velocity, temperature, radius, and degree

of crystallinity; z is the axial jet coordinate; szz and srr are

the axial and radial normal stresses in the jet; and finally, hand T1 are the convective heat transfer coefficient and the

bulk cooling air temperature, respectively.

The various terms in the energy equation have been dis-

cussed in detail previously,13–15 but the radial convection

(last term on the RHS) is of primary interest to this work. It

stems from the jet surface heat flux boundary condition and

enters the conservation equation due to thin filament approx-

imation used in one-dimensionalization of the problem. The

heat transfer coefficient is typically obtained from empirical

correlations, relating the system geometry and fluid flow

characteristics to the thermal boundary layer formation. Our

previous model included a correlation proposed by Kase and

Matsuo (KM),18,19 which is widely used in conventional

melt spinning literature and produces good agreement with

experimentally observed thermal profiles.8–12 The KM corre-

lation is based on the empirical formula for longitudinal

external flow along a cylinder:

NuD ¼ 0:42 Re1=3; (2)

where NuD is the Nusselt number (hDjet=kair) and Re is the

Reynolds number (VairDjet=�air). Djet is the polymer jet di-

ameter, Vair is the axial air velocity, kair is the air thermal

conductivity, and �air is the air kinematic viscosity. Substi-

tuting, rearranging, and accounting for the jet axial velocity,

V, one obtains:

h ¼ 0:388 kairVair � Vj jvairAjet

� �1=3

; (3)

where Ajet is the jet cross-sectional area. Kase and Matsuo

introduced an additional correction factor to account for

transverse air flow, Vc,air, to obtain their final correlation:

h ¼ 0:388 kairVair � Vj jvairAjet

� �1=3

1þ 8Vc;air

Vair

� �2( )1=6

: (4)

The air properties are evaluated at film temperature, Tfilm,

defined as the arithmetic average of the jet and bulk air tem-

peratures. For dry air, the following correlations18 (in SIunits) can be used in the range 300 K<Tfilm< 650 K, with

an expected error of less than 2%:

kair ¼ 1:8786� 10�4T0:866film ; (5)

�air ¼ 4:29059� 10�9T

5=2film

Tfilm þ 120: (6)

B. Conventional heat transfer enhancementmechanisms

Thermal images collected in the present study have

revealed that the quenching heat transfer rate in electrospin-

ning is significantly higher, �10-fold, than predicted by KM

correlation. Therefore, we consider a number of alternative

heat transfer mechanisms and enhancements.

FIG. 1. (Color online) Melt electrospinning setup schematic.

TABLE I. Experimental conditions.

Case Polymer

Nozzle temperature,

Tnozzle (K)

Flow rate,

Q� 10�10 (m3=s)

1 N6 543 1.17

2 N6 563 2.5

3 N6 583 5.0

4 PP 513 5.0

5 PLA 483 1.67

073102-2 Zhmayev, Cho, and Joo Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

1. Natural convection effects

Natural convection is typically neglected in mechanical

melt spinning in favor of forced convection due to relatively

high jet velocities. In the case of melt electrospinning, partic-

ularly near the spinneret, Re for air surrounding the jet can

often be on the order of 1 or less, and thus, natural convec-

tion may be more important than assumed by KM. The fol-

lowing correlation can be used to estimate natural

convection in horizontal cylinder geometry:20

NuD ¼hD

kair¼ 0:36þ 0:518Ra

1=4D

0:559= Prairð Þ9=16þ1� �4=9

; (7)

RaD ¼gD3

vairaair

T � T1T1

� �; (8)

where RaD is the Rayleigh number, g is acceleration due to

gravity, and Prair and aair are the Prandtl number and thermal

diffusivity of air, evaluated at film temperature. The air prop-

erties can be calculated using correlations (5) and (6), with

heat capacity being essentially constant at 1009 J=kg-K.

2. Radiation effects

Radiation heat transfer is typically negligible below

�600 K compared to convection. Even though most of the

jet is at lower temperatures, radiation close to the spinneret,

especially in electrospinning of polymers with a high melting

point, such as N6, may be significant. Radiation can be esti-

mated using gray body formulation:21

Qrad ¼ h T � T1ð Þ ¼ reR T4 � T41

� �; (9)

where Qrad is the radiative heat flux, r is the Stefan-

Boltzmann constant, and eR is the polymer jet emissivity.

3. Humidity effects

It has been observed experimentally that environmental

humidity can have significant effects on spinnability. Hygro-

scopic polymers, e.g., N6, are especially sensitive. They

have to be thoroughly dried and carefully handled prior to

spinning, and environmental humidity has to be controlled.

Even for non-hygroscopic polymers, e.g., PLA, it was

observed that spinning conditions have to be adjusted

depending on air humidity. Literature suggests that air hu-

midity can result in as much as 20% heat transfer enhance-

ment.22,23 This enhancement was modeled as a constant

multiplier to the KM correlation.

C. Electrohydrodynamic (EHD) heat transferenhancement

The conventional heat transfer mechanisms discussed in

Sec. III B are not expected to produce the experimentally

evident 10-fold heat transfer enhancement relative to the

KM correlation. The largest enhancement mechanism

was found in the heat exchange and electronic circuit litera-

ture24–33 and is due to electrohydrodynamic (EHD) effects. It

has been documented that strong electric fields can result in

significant corona currents, which disturb the thermal bound-

ary layer and can enhance the heat transfer rate as much as

60-fold in some applications. When air is used as the

quenching fluid, an order of magnitude heat transfer

enhancement is typical.24–27,32

1. Mechanism of the EHD-driven air flow

EHD heat transfer enhancement has not been previously

documented in electrospinning, but a recent publication34

and our experimental observations did note stable corona

TABLE II. Polymer material properties used in modeling.

Properties=Parameters N6 (Ref. 15) PP (Ref. 16) PLA (Ref. 13)

Melt density, q (kg=m3) 1000 770 1240

Heat capacity, Cp (J=kg K) 2553 1925 1800

Thermal conductivity, kT (W=m K) 0.245 0.12 0.2

Dielectric constant ratio, e=e0 3.2 2.25 3.1

Surface tension, c (N=m) 0.030 0.0221 0.0435

Stress relaxation time, k (s) 0.025 @ 523 K 0.035 @ 453 K 0.1 @ 453 K

Giesekus mobility factor, a 0.04 0.4 0.015

Activation energy of flow, DH=Rig (K) 6600 2350 9060

Zero-shear-rate viscosity, g0 (Pa s) 193 @ 523 K 905 @ 453 K 1320 @ 453 K

Ratio of solvent to zero-shear-rate viscosity, b 0.3 0.01 0.35

Emissivity, eR 0.56 @ 445 K 0.68 @ 443 K 0.68 @ 425 K

FIG. 2. Coronating electrode current versus applied voltage.

073102-3 Electrohydrodynamic quenching Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

formation. In contrast with electrically conducting polymer

solutions commonly studied in electrospinning literature,

polymer melts are highly insulating, and when a high voltage

is applied to a needle-plate geometry separated by dielectrics

(in this case air and polymer melt), a corona field is often

produced. This condition is obtained when the applied poten-

tial exceeds the corona onset, but is below the air gap break-

down voltage. During the corona discharge, the local electric

field near the needle tip exceeds the breakdown electric field

strength in air, E0¼ 3.2� 106 V=m, and as a result, air mole-

cules become ionized. This region is referred to as the ioni-

zation zone, and it contains ions of both polarities. Further

away from the needle, where the local electric field decreases

below E0, no further ionization occurs. However, in this

drifting zone, also known as the space charge plume, ions of

the same polarity as the needle are transported from the nee-

dle to the plate due to action of the electrostatic forces.

Along their flight path, the ions can exchange momentum

with and accelerate neutral air molecules. In the present

case, high positive voltage is applied to the needle, so a posi-

tive corona is formed, and thus, positive air ions are observed

in the drifting zone. Schematic of this process is shown in

Figure 3 and further detailed discussions can be found in

literature.17,24–33,35

The set of governing equations for this process has been

reported previously17,25 and is as follows:

Conservation of momentum :

qairVair � rVair ¼ �rpþ lairr2Vair þ qE; (10)

Mass continuity : r � Vair ¼ 0; (11)

Ionic current density : J ¼ lEEqþ Vairq� DErq; (12)

Charge conservation : r � J ¼ 0; (13)

Poisson0s equation : r2U ¼ �q=e; (14)

Electric Field : E ¼ �rU; (15)

where qair and lair are the air density and dynamic viscosity,

p is the pressure, q is the space charge density, E is the elec-

tric field, J is the electric current density, lE is the air ion

mobility (2.0� 10�4 m2=V-s for positive air ions31), DE is

the ion diffusivity coefficient, U is the electric potential, and

e is the permittivity of air. The last term in the momentum

equation, Eq. (10), is the Coulombic force exerted on the air

by the action of space charge in the electric field.

Solution of this set of coupled equations requires

specialized numerical techniques discussed in the litera-

ture.17,25,27,29,31 In this work, we utilize the findings from

various numerical studies to simplify the system and obtain

an approximate analytical expression for the air velocity

along the polymer melt jet and consequently determine the

EHD heat transfer enhancement.

2. Axial velocity approximate solution

In this study, we are particularly concerned with the

axial air velocity along the spinline axis, which disturbs the

thermal boundary layer in the region from the needle tip up

to �35 needle radii (R0) from the needle. To obtain an ap-

proximate solution, we utilize the findings of a numerical

investigation for a similar geometry.17

First let us consider the conservation of momentum, Eq.

(10). In the numerical study, the air flow streamlines in the

vicinity of the spinline axis were found to be nearly parallel

to the axis. Furthermore, as discussed elsewhere,35 in the

boundary layer close to the polymer jet, we can neglect the

pressure perturbation and the viscous term due to large air

Re (expected to be on the order of 1000). The momentum

balance is then reduced to:

qairVz;airdVz;air

dz¼ qEz: (16)

To determine the Coulombic force term, we turn to the cur-

rent and charge conservation Eqs. (12) and (13). From the

numerical investigation, the space charge plume in air takes

on a nearly conical shape, with the space charge density

decreasing with distance from the needle due to Coulomb

repulsion. As a first approximation, it could be assumed that

the charge density is inversely proportional to the cross-sec-

tional area of the plume cone, which in turn would be

approximated as p(R0þz)2. Since current transport is limited

to this conical plume region, the axial current flux averaged

over the transverse direction, Jz,avg, is then:

Jz;avg zð Þ ¼ I

p R0 þ zð Þ2; (17)

where I is the total current emitted from the coronating nee-

dle tip and R0

is the needle radius. Using this average flux to

determine the Coulombic acceleration would be an underes-

timation. From experimental measurements,36 it is evident

that the charge density and, therefore, the current flux are at

their maximum along the spinline axis and decay rapidly in

the transverse direction. Based on the experimental data, we

propose the following radial current flux, Jr, profile:

Jr r; hð Þ ¼ Ccos2 h

r2; (18)

where h is the inclination angle with respect to the spinline

axis and C is the normalization constant, which can be deter-

mined by spherical averaging of 18 and comparing to 17 toFIG. 3. (Color online) Schematic of EHD-driven air flow in melt

electrospinning.

073102-4 Zhmayev, Cho, and Joo Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

find C¼ 3 I = p. We then obtain a better approximation for

the axial current flux along the spinline:

Jz zð Þ ¼ 3I

p R0 þ zð Þ2: (19)

Finally, based on previous studies in the literature31,35 and

our estimations on the magnitudes of various terms in

Eq. (12), the conduction term (lEEq) dominates the convec-

tive and diffusive fluxes (last two terms) in the region near

the needle, which is of interest to this study. Further from the

needle, the air velocity can increase sufficiently to make the

convection term comparable to conduction. However, at this

point, the space charge density would have diminished sig-

nificantly, thus, causing the current flux and consequently

the Coulombic acceleration term in the momentum balance

to be low.

With the above assumptions and approximations, we

arrive at the axial momentum balance near the spinline axis:

qairVz;airdVz;air

dz¼ 3I

lEp R0 þ zð Þ2; (20)

and after applying the boundary condition Vz,air(0)¼ 0, we

get:

Vz;airðzÞ ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi6I

qairlEpR0

1� 1

1þ z=R0

� �s: (21)

This approach was found to provide the correct qualitative

behavior of the air velocity along the spinline, but some of

the simplifying assumptions may limit its accuracy. Further

refinements to this model are discussed in Sec. III C 3.

3. EHD model refinements

In the preceding discussion, the transverse components

of current and velocity, as well as the pressure perturbations

were eliminated. These restrictions may limit the accuracy of

the calculated axial air velocity and also conceal the rapid

decay of the velocity further from the spinline. These

restrictions can be removed by solving the conservation of

momentum in the stream function form. However, to obtain

a non-trivial solution, the current flux must be known and it

must have both r- and h- components. No analytical solution

for the current flux in this problem has been reported to date,

and we have not been able to find a suitable approximation

that satisfies all of the governing Eqs. (13)–(15). Therefore,

at present, we must rely on numerical methods or crude

approximations to elucidate the EHD effects.

IV. RESULTS AND DISCUSSION

The calculated EHD-driven air velocity profiles for N6

are compared to the jet velocities in Figure 4(a). Due to co-

rona discharge in the vicinity of the needle, the air in the

drifting zone is rapidly accelerated. It is apparent that the

resulting air velocities are three orders of magnitude higher

than the jet velocities. Subsequently to the short initial tran-

sient (z=R0 < 3), the driving electric field rapidly decreases

and the air velocity approaches its asymptotic value. We

should note that eventually the velocity should decrease due

to viscous forces, but we have assumed that these effects are

not significant in the region of interest, z=R0 < 35.

FIG. 4. (Color online) Calculations for N6 met electrospinning: (a) EHD-

driven air velocity profiles compared to the jet velocities, (b) KM local heat

transfer coefficient profiles, and (c) heat transfer enhancement due to natural

convection, radiation, and EHD effects.

073102-5 Electrohydrodynamic quenching Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

The KM heat transfer coefficient profiles for N6 are

shown in Figure 4(b). The rapid initial increase in heat trans-

fer is due to high jet attenuation, and the asymptotic values

are reached when the jet profiles become nearly cylindrical.

The higher polymer flow rates for higher nozzle temperature

cases produce higher jet velocities and thus higher heat trans-

fer coefficients.

Observing the various heat transfer enhancement ratios,

Figure 4(c), it can be concluded that radiation effects are in

fact negligible, and natural convection provides only a slight

enhancement very near the spinneret. Interestingly, the natu-

ral convection heat transfer coefficient in its asymptotic

region is close to the KM correlation predictions. The 20%

enhancement due to air humidity (not shown) is clearly not

sufficient to match experimental observations. Finally, the

EHD effects produced the expected pronounced enhance-

ment, consistent with experimental observations. While only

N6 cases are shown, similar results were obtained for PP and

PLA.

As discussed above, the KM model predicts significantly

slower jet cooling than observed experimentally, illustrated

for the case of N6 at nozzle temperature of 543 K in Figure

5. Incorporation of the proposed EHD mechanism allowed

for the correction of the thermal profile prediction, while

other possible heat transfer enhancement mechanisms failed

to match experimental observations. Since the emission cur-

rent for the present case (221 lA) was obtained by extrapola-

tion of numerical data, in Figure 5 we also show results for

other currents to demonstrate that the exact value of the

emission current is not required for these calculations to pro-

vide suitable agreement with experiments. For clarity, exper-

imental error bars are not included in the figure, but based on

camera resolution, accuracy of emissivity estimation, and

image variability, the associated experimental error varies

from 62% at the nozzle to 610% as room temperature is

reached. The major reasons for the increased error at lower

temperature are the influence of the background temperature

and the emissivity estimation based on the high temperature

thick jet condition.

Experimental jet radius and thermal profiles along the

spinning axis and the model results with EHD enhancement

for a number of polymer melts are presented in Figure 6.

Using the present model, excellent agreement with all of the

FIG. 5. (Color online) Experimental jet temperature profile for N6 (at noz-

zle temperature 543 K) compared to KM predictions, radiation and natural

convection enhancements, and the present EHD model.

FIG. 6. (Color online) Experimental jet radius and temperature profiles for

(a) N6, (b) PP, and (c) PLA compared to model predictions. Notes: for N6,

the images shown in the insets are for 563 K nozzle temperature; in all ther-

mal images, the needle edges have been outlined in post-processing.

073102-6 Zhmayev, Cho, and Joo Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

investigated cases was achieved. The discrepancy in the ther-

mal profiles for the N6 583 K case should be attributed to

low thermal image quality in the thin jet region (most likely

due to the jet veering away from the camera focal plane,

which is commonly observed at higher temperatures). While

only the results with EHD enhancement are shown in Figure

6, the other discussed heat transfer modes were also calcu-

lated. As demonstrated above, the other mechanisms result

in much lower heat transfer coefficients, and thus signifi-

cantly different thermal profiles. The agreement of the pres-

ent model with the experimental data indicates that the

proposed EHD heat transfer enhancement mechanism in

electrospinning is plausible, and most of the assumptions

used in the model derivation are reasonable in the region of

importance. Further investigations on the assumptions of the

uniform current flux and negligible pressure perturbation are

warranted.

Based on the performed analysis and discussions, the

undesirable EHD quenching can be minimized=eliminated

using the following techniques. First, forced convection of

heated ambient air would reduce the jet cooling by not only

minimizing the temperature differential between the jet and

the air, but also by perturbing the EHD quenching mecha-

nism.33 Second, applying high potential to the collector,

rather than the spinneret, would allow for elimination of co-

rona discharge, while preserving the electric field profile.

These modifications have been implemented in a different

study37 and were shown to be effective at preventing the

undesirable quenching, and consequently resulting in pro-

duction of smaller fibers from polymer melt electrospinning.

V. CONCLUSIONS

In summary, IR thermal measurements have revealed

enhanced air quenching of polymer melt jets in electrospin-

ning. We have postulated that EHD effects are the root cause

of this quenching and developed a correlation to account for

this enhancement. The correlation was implemented in a

comprehensive model for polymer melt electrospinning and

validated by excellent agreement with experimental observa-

tions. Based on this analysis, we concluded that addition of

forced hot air convection and application of high potential to

the collector rather than needle can be used to minimize/

eliminate the undesirable jet quenching, which will in turn

allow for production of smaller fibers from polymer melt

electrospinning.

ACKNOWLEDGMENTS

The authors would like to thank the National Science

Foundation for funding this work through CAREER Award,

Grant No. CTS-0448270, and E.I. du Pont de Nemours and

Company for funding through DuPont Young Professor

Grant to YLJ.

1H. Fong and D. H. Reneker, “Electrospinning and the formation of nano-

fibers,” in Structure Formation in Polymeric Fibers, edited by D. R. Salem

(Hanser Gardner Publications, Munich, 2001).2Y. Dzenis, “Spinning continuous fibers for nanotechnology,” Science 304,

1917 (2004).

3B. D. Li and Y. Xia, “Electrospinning of nanofibers: Reinventing the

wheel?,” Adv. Mater. 16, 1151 (2004).4L. Larrondo and S. J. Manley, “Electrostatic fiber spinning from polymer

melts. I. Experimental observations on fiber formation and properties, II.

Examination of the flow field in an electrically driven jet, III. Electrostatic

deformation of a pendant drop of polymer melt,” J. Polym. Sci., Polym.

Phys. Ed. 19, 909 (1981).5J. S. Kim and D. S. Lee, “Thermal properties of electrospun polyesters,”

Polym. J. 32, 616 (2000).6J. Lyons, C. Li, and F. Ko, “Melt-electrospinning part I: Processing param-

eters and geometric properties,” Polymer 45, 7597 (2004).7H. Zhou, T. B. Green, and Y. L. Joo, “The thermal effects on electrospin-

ning of polylactic acid melts,” Polymer 47, 7497 (2006).8A. Ziabicki, Fundamentals of Fiber Formation: The Science of Fiber Spin-ning and Drawing (Wiley, New York, 1976).

9M. M. Denn, “Fibre Spinning,” in Computational Analysis of PolymerProcessing, edited by J. R. A. Pearson and S. M. Richardson (Applied Sci-

ence Publisher, London, 1983).10B. P. Huynh and R. I. Tanner, “Study of the non-isothermal glass fibre

drawing process,” Rheol. Acta 22, 482 (1983).11G. M. Henson, D. Cao, and S. E. Bechtel, “A thin-filament melt spinning

model with radial resolution of temperature and stress,” J. Rheol. 42, 329

(1998).12Y. L. Joo, J. Sun, M. D. Smith, R. C. Armstrong, R. A. Brown, and R. A.

Ross, “2-D numerical analysis of nonisothermal melt spinning with and

without phase transition,” J. Non-Newtonian Fluid Mech. 102, 37 (2002).13E. Zhmayev, H. Zhou, and Y. L. Joo, “Modeling of non-isothermal poly-

mer jets in melt electrospinning,” J. Non-Newtonian Fluid Mech. 153, 95

(2008).14E. Zhmayev and Y. L. Joo, “Modeling of crystallizing polymer melts in

electrospinning,” AIP Conf. Proc. 1027, 9 (2008).15E. Zhmayev, D. Cho, and Y. L. Joo, “Modeling of melt electrospinning for

semi-crystalline polymers,” Polymer 51, 274 (2010).16W. H. Kohler and A. J. McHugh, “Sensitivity analysis of low-speed melt

spinning of isotactic polypropylene,” Chem. Eng. Sci. 62, 2690 (2007).17L. Zhao and K. Adamiak, “EHD flow in air produced by electric corona

discharge in pin-plate configuration,” J. Electrostat. 63, 337 (2005).18S. Kase and T. Matsuo, “Studies on melt spinning. I. Fundamental equa-

tions on the dynamics of melt spinning,” J. Polym. Sci. Part A 3, 2541

(1965).19S. Kase and T. Matsuo, “Studies on melt spinning. II. Steady-state and

transient solutions of fundamental equations compared with experimental

results,” J. Appl. Polym. Sci. 11, 251 (1967).20S. W. Churchill and H. H. S. Chu, “Correlating equations for laminar and

turbulent free convection from a horizontal cylinder,” Int. J. Heat Mass

Transfer 18, 1049 (1975).21J. H. Lienhard IV and J. H. Lienhard V, A Heat Transfer Textbook (Phlo-

giston, Cambridge, MA, 2002).22J. Zhang, A. Gupta, and J. Baker, “Effect of relative humidity on the pre-

diction of natural convection heat transfer coefficients,” Heat Transfer

Eng. 28, 335 (2007).23M. Still, H. Venzke, F. Durst, and A. Melling, “Influence of humidity on

the convective heat transfer from small cylinders,” Exp. Fluids 24, 141

(1998).24D. B. Go, R. A. Maturana, T. S. Fisher, and S. V. Garimella,

“Enhancement of external forced convection by ionic wind,” Int. J. Heat

Mass Transfer 51, 6047 (2008).25A. Bandyopadhyay, S. Sen, A. Sarkar, and S. Bhattacharyya,

“Investigation of forced convective heat transfer enhancement in the pres-

ence of an electric field – a finite element analysis,” Int. J. Numer. Meth-

ods Heat Fluid Flow 14, 264 (2004).26S.-A. O. Ahmedou and M. Havet, “Effect of process parameters on the

EHD airflow,” J. Electrostat. 67, 222 (2009).27M. Molki and P. Damronglerd, “Electrohydrodynamic enhancement of

heat transfer for developing air flow in square ducts,” Heat Transfer Eng.

27, 35 (2006).28C. C. Pascual, J. H. Stromberger, S. M. Jeter, and S. I. Abdel-Khalik, “An

empirical correlation for electrohydrodynamic enhancement of natural

convection,” Int. J. Heat Mass Transfer 43, 1965 (2000).29Z. Djuric, W. Balachandran, and C. W. Wilson, “Electrical field and space

charge modeling in a viscous fluid flow in a nozzle,” J. Phys. D: Appl.

Phys. 31, 2132 (1998).30M. Goldman, A. Goldman, and R. S. Sigmond, “The corona discharge, its

properties and specific uses,” Pure Appl. Chem. 57, 1353 (1985).

073102-7 Electrohydrodynamic quenching Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

31N. E. Jewell-Larsen, C. P. Hsu, I. A. Krichtafovitch, S. W. Montgomery, J.

T. Dibene, II, and A. V. Mamishev, “CFD analysis of electrostatic fluid

accelerators for forced convection cooling,” IEEE Trans. Dielectr. Electr.

Insul. 15, 1745 (2008).32S. Laohalertdecha, P. Naphon, and S. Wongwises, “A review of electrohy-

drodynamic enhancement of heat transfer,” Renewable Sustainable Energy

Rev. 11, 858 (2007).33S. R. Parsons, “The effect of corona current on the cooling of a hot wire,”

Phys. Rev. 33, 75 (1929).

34S. Tripatanasuwan and D. Reneker, “Corona discharge from electrospin-

ning jet of poly(ethylene oxide) solution,” Polymer 50, 1835 (2009).35P. Atten, B. Malraison, and M. Zahn, “Electrohydrodynamic plumes in

point-plane geometry,” IEEE Trans. Dielectr. Electr. Insul. 4, 710 (1997).36P. Atten, K. Adamiak, B. Khaddour, and J.-L. Coulomb, “Simulation of

corona discharge in configurations with a sharp electrode,” J. Optoelec-

tron. Adv. Mater. 6, 1023 (2004).37E. Zhmayev, D. Cho, and Y. L. Joo, “Nanofibers from gas-assisted poly-

mer melt electrospinning,” Polymer 51, 4140 (2010).

073102-8 Zhmayev, Cho, and Joo Phys. Fluids 23, 073102 (2011)

Downloaded 08 May 2012 to 132.236.27.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions