15
Rheol Acta (2012) 51:909–923 DOI 10.1007/s00397-012-0649-3 ORIGINAL CONTRIBUTION Determination of axial forces during the capillary breakup of liquid filaments – the tilted CaBER method Dirk Sachsenheimer · Bernhard Hochstein · Hans Buggisch · Norbert Willenbacher Received: 24 February 2012 / Revised: 11 July 2012 / Accepted: 26 July 2012 / Published online: 28 August 2012 © Springer-Verlag 2012 Abstract The capillary breakup extensional rheometry (CaBER) is a versatile method to characterize the elon- gational behavior of low-viscosity fluids. Commonly, data evaluation is based on the assumption of zero normal stress in axial direction (σ zz = 0). In this pa- per, we present a simple method to determine the axial force using a CaBER device rotated by 90 and analyzing the deflection of the filament due to grav- ity. Forces in the range of 0.1–1,000 μN could be as- sessed. Our study includes experimental investigations of Newtonian fructose solutions and silicon oil mixtures (viscosity range, 0.9–60 Pa s) and weakly viscoelastic polyethylene oxide (PEO, M w = 10 6 g/mol) solutions covering a concentration range from c c(critical overlap concentration) up to c > c e (entanglement con- centration). Papageorgiou’s solution for the stress ratio σ zz /σ rr in Newtonian fluids during capillary thinning is experimentally confirmed, but the widely accepted assumption of vanishing axial stress in weakly viscoelas- tic fluids is not fulfilled for PEO solutions, if c e is exceeded. Keywords CaBER · Elongational viscosity · Uniaxial extension · Force measurement D. Sachsenheimer (B ) · B. Hochstein · H. Buggisch · N. Willenbacher Karlsruhe Institute of Technology (KIT), Institute for Mechanical Process Engineering and Mechanics, Group Applied Mechanics (AME), Gotthard-Franz-Straße 3, 76131 Karlsruhe, Germany e-mail: [email protected] Introduction General remarks Many industrial applications and processes such as coating (Fernando et al. 1989, 2000), spraying (Dexter 1996; Prud’homme et al. 2005) (including mist for- mation (James et al. 2003) and its prevention (Chao et al. 1984)), inkjet printing (Agarwal and Gupta 2002; Han et al. 2004; Vadillo et al. 2010) or fiber spinning (McKay et al. 1978) include flow kinematics with large elements of elongational flow. Due to this high technical relevance, the correlation between rhe- ological properties against elongational deformation and fluid behavior in processes is a major subject of research. Technically relevant liquids are often complex multi- component systems with special flow properties ad- justed by adding small amounts of rheological modifier or thickener. A great number of these additional materials are commercially available, e.g., biopoly- mers (polysaccharides) like xanthan gum, starch, car- rageenan, or especially cellulose derivatives, inorganic substances like silica or water-swellable clay, or simply synthetic polymers like polyacrylates, polyvinylpyrili- done, or polyethylene oxide (PEO) (Braun and Rosen 2000). Therefore, understanding the elongational flow properties of such viscoelastic polymer solutions is of fundamental importance in process optimization and product development. Unfortunately, measuring the elongational viscosity of low viscosity fluids is still a very challenging task, resulting in only a few investiga- tions which correlate the elongational behavior of com- plex fluids with their application properties (Meadows et al. 1995; Kennedy et al. 1995; Ng et al. 1996; Solomon

Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923DOI 10.1007/s00397-012-0649-3

ORIGINAL CONTRIBUTION

Determination of axial forces during the capillary breakupof liquid filaments – the tilted CaBER method

Dirk Sachsenheimer · Bernhard Hochstein ·Hans Buggisch · Norbert Willenbacher

Received: 24 February 2012 / Revised: 11 July 2012 / Accepted: 26 July 2012 / Published online: 28 August 2012© Springer-Verlag 2012

Abstract The capillary breakup extensional rheometry(CaBER) is a versatile method to characterize the elon-gational behavior of low-viscosity fluids. Commonly,data evaluation is based on the assumption of zeronormal stress in axial direction (σzz = 0). In this pa-per, we present a simple method to determine theaxial force using a CaBER device rotated by 90◦ andanalyzing the deflection of the filament due to grav-ity. Forces in the range of 0.1–1,000 μN could be as-sessed. Our study includes experimental investigationsof Newtonian fructose solutions and silicon oil mixtures(viscosity range, 0.9–60 Pa s) and weakly viscoelasticpolyethylene oxide (PEO, Mw = 106 g/mol) solutionscovering a concentration range from c ≈ c∗ (criticaloverlap concentration) up to c > ce (entanglement con-centration). Papageorgiou’s solution for the stress ratioσzz/σrr in Newtonian fluids during capillary thinningis experimentally confirmed, but the widely acceptedassumption of vanishing axial stress in weakly viscoelas-tic fluids is not fulfilled for PEO solutions, if ce isexceeded.

Keywords CaBER · Elongational viscosity ·Uniaxial extension · Force measurement

D. Sachsenheimer (B) · B. Hochstein ·H. Buggisch · N. WillenbacherKarlsruhe Institute of Technology (KIT),Institute for Mechanical Process Engineeringand Mechanics, Group Applied Mechanics (AME),Gotthard-Franz-Straße 3, 76131 Karlsruhe, Germanye-mail: [email protected]

Introduction

General remarks

Many industrial applications and processes such ascoating (Fernando et al. 1989, 2000), spraying (Dexter1996; Prud’homme et al. 2005) (including mist for-mation (James et al. 2003) and its prevention (Chaoet al. 1984)), inkjet printing (Agarwal and Gupta2002; Han et al. 2004; Vadillo et al. 2010) or fiberspinning (McKay et al. 1978) include flow kinematicswith large elements of elongational flow. Due to thishigh technical relevance, the correlation between rhe-ological properties against elongational deformationand fluid behavior in processes is a major subject ofresearch.

Technically relevant liquids are often complex multi-component systems with special flow properties ad-justed by adding small amounts of rheological modifieror thickener. A great number of these additionalmaterials are commercially available, e.g., biopoly-mers (polysaccharides) like xanthan gum, starch, car-rageenan, or especially cellulose derivatives, inorganicsubstances like silica or water-swellable clay, or simplysynthetic polymers like polyacrylates, polyvinylpyrili-done, or polyethylene oxide (PEO) (Braun and Rosen2000). Therefore, understanding the elongational flowproperties of such viscoelastic polymer solutions is offundamental importance in process optimization andproduct development. Unfortunately, measuring theelongational viscosity of low viscosity fluids is still avery challenging task, resulting in only a few investiga-tions which correlate the elongational behavior of com-plex fluids with their application properties (Meadowset al. 1995; Kennedy et al. 1995; Ng et al. 1996; Solomon

Page 2: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

910 Rheol Acta (2012) 51:909–923

and Muller 1996; Tan et al. 2000; Stelter et al. 2002; Ploget al. 2005).

Nevertheless, a simple and versatile method forthe characterization of low-viscosity fluids has beensuggested more than 20 years ago (Bazilevsky et al.1990; Entov and Hinch 1997; Bazilevsky et al. 2001),this so called capillary breakup extensional rheometer(CaBER) is even commercially available now. In thisexperiment, an instable liquid filament is created byapplying a step strain, and the diameter of this liquidbridge is monitored as a function of time. In contrastto other techniques, CaBER allows for large Henckystrains which are of great significance to industrial prac-tice. Unfortunately, evaluating the elongational viscos-ity is not trivial for CaBER experiments since no axialforce is measured. Accordingly, apparent elongationalviscosities are often calculated for CaBER.

In this paper, we introduce the tilted CaBER methodas a simple and accurate way to determine the axialforce F(t) and the elongational rate ε̇ during filamentthinning at the same time by only analyzing videoimages. The conventional CaBER device is rotatedby 90◦, and the bending of the liquid filament due togravity is recorded using a high-speed camera. First,we describe the principles of calculating the (appar-ent) elongational viscosity from CaBER measurementsand discuss the typical decrease of the diameter forNewtonian and viscoelastic fluids. Then, we describehow to calculate the force from the deflection of a fluidfilament. After this, we present the experimental setupand give a short overview of the samples used, includingpreparation and characterization. Following with theexperimental part, we verify the tilted CaBER methodusing Newtonian fluids and apply it to non-NewtonianPEO solutions. Finally, concluding remarks are given.

Force balance for a straight vertical cylindrical thread

The axial force F in a cylindrical filament with axial ori-entation into the direction of gravity (here, z-direction)is assumed to be independent of the z-position, but Fmay depend on time. Taking into account the surfacetension �, the normal stresses σzz and σrr, but neglect-ing gravity and inertia effects (see Fig. 1), the total forcebalances in z-direction reads as follows:

π

4σzz D2 + π�D = F ⇒ σzz = 4 (F − π�D)

π D2 . (1)

McKinley and Tripathi (2000) used the ratio X be-tween the true axial force F in the filament and thevalue resulting from the assumption σzz = 0 in order to

Fig. 1 Cut filament (hatched areas) with normal stresses σzz andσrr

quantify the influence of the axial normal stress σzz.

X = FFσzz=0

= Fπ�D

(2)

For an infinitesimal volume element with length dzand diameter D, the force balance in r-direction readsas follows:

σrr D dz + 2� dz = 0 ⇒ σrr = −2�

D. (3)

Calculation of the elongational viscosity

The elongational viscosity ηe for uniaxial elongationalflows like, e.g., in CaBER experiments is given bySchümmer and Tebel (1983):

ηe = σzz − σrr

ε̇. (4)

Insertion of the axial normal stress σzz (Eq. 1) andthe radial normal stress σrr (Eq. 3) into Eq. 4 results inthe following expression for the true elongational vis-cosity valid for cylindrical or at least slender filaments:

ηe = 4F − 2π�Dπ D2ε̇

. (5)

Using the definition of the elongation rate

ε̇ = − 2D

dDdt

, (6)

equation 5 yields

ηe = �

dD/dt− 2F

π D dD/dt. (7)

For the sake of completeness, expressions for theradial normal stress and the elongational viscosity incase of non-cylindrical filaments are given in Appendix.

Page 3: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 911

It is obvious from Eq. 7 that for calculating ηe, thediameter D and the force F are needed. Therefore,CaBER experiments with simultaneous force measure-ments allow for the determination of the true elon-gational viscosity for any fluid without assumption ofspecial constitutive equations.

But the evaluation of Eq. 7 is not trivial for CaBERexperiments since no axial force measurement is in-cluded, and, hence, the normal stress σzz is not known.Instead, ηe is often evaluated using the assumptionσzz = 0. Then an apparent elongational viscosity forCaBER experiments is obtained (Anna and McKinley2001):

ηe,app = − �

dD/dt. (8)

So far, the σzz = 0 assumption has not been validatedexperimentally for the CaBER experiment. Deviationsobviously occur for the Newtonian case (Liang andMackley 1994; Kolte and Szabo 1999; McKinley andTripathi 2000) and also shown in numerical simulationsof the CaBER experiment using different viscoelas-tic or viscoplastic constitutive equations (Clasen et al.2006a; Webster et al. 2008; Alexandrou et al. 2009).

Therefore, fluid characterization based on Eq. 8might be helpful in comparative studies employed, e.g.,for product development purposes. But especially forthe determination of the true elongational viscosityfrom CaBER experiments, it is mandatory to measurethe time-dependent axial force F during filament thin-ning accurately. The magnitude of this force can beroughly estimated from the force balance in axial direc-tion (Eq. 1) using the σzz = 0 assumption. Typically, �

is in the range of 20–70 mN/m, and the filament diam-eter decays from 1 mm to about 10 μm. This corre-sponds to a force range 0.5 μN < F(t) < 220 μN. Thefirst attempt to implement a force transducer into aCaBER device was done by Klein et al. (2009). Theymounted a commercial quartz load cell to the fixedbottom plate of their apparatus. Due to oversamplingof the force signal, a nominal sensitivity of 50 μN wasreached, but calibration was only done in the forcerange larger than 2,000 μN. Thus, reliable force detec-tion during capillary thinning of fluid filaments was notpossible.

Time evolution of the filament diameterfor Newtonian and weakly viscoelastic fluids

The time evolution of the diameter during capillarythinning is controlled by a balance of capillary andviscous or viscoelastic forces. Different characteristicin diameter vs. time curves are observed for different

types of fluids, e.g., Bingham plastic, power law, andNewtonian or viscoelastic fluids.

The midpoint diameter Dmid of a Newtonian fluidin a CaBER experiment decreases linearly with time taccording to Papageorgiou (1995) and McKinley andTripathi (2000):

Dmid (t) = D1 − ��

ηst (9)

where D1 is the initial diameter of the liquid bridge attime t = 0 at which linear thinning of the filament setsin, � is the surface tension, � is a constant numericalfactor, and ηs is the shear viscosity. The local forcebalance (σzz = 0) for a thinning liquid yields � = 0.333if only the surface tension is considered. Papageorgiou(1995) calculated the numerical factor to � = 0.1418 incase of negligible inertia (Reynolds number Re → 0).Considering inertia (Re > 0), Eggers (1993, 1997) andBrenner et al. (1996) estimated the numerical factor to� = 0.0608 and showed that this solution is valid nearthe filament break up.

McKinley and Tripathi (2000) confirmed the nu-merical results of Papageorgiou (1995) (� = 0.1418)experimentally using glycerol samples and related thefactor � to the force ratio X (Eq. 2).

X = 3� + 12

(10)

The experimental validated � = 0.1418 value cor-responds to X = 0.713. This indicates that the axialforce is not only given by the surface tension, andan additional axial normal stress must be present forNewtonian liquid filaments.

In contrast to Newtonian liquids, weakly elastic poly-mer solutions with concentrations c < ce form cylin-drical filaments, and their diameter decreases expo-nentially with time in CaBER experiments (Bazilevskyet al. 1990; Entov and Hinch 1997; Anna and McKinley2001; Arnolds et al. 2010).

D (t) = D1

(GD1

)1/3

exp(

− t3λe

)(11)

where D1 is the initial diameter of the filament atthe beginning of the exponential decrease, G is theelastic modulus, and λe is the elongational relaxationtime, which is related to the constant elongation rateε̇ = 2/ (3λe) (see also Eq. 6).

Exponential thinning for viscoelastic fluids can alsoresult if an exponentially increasing normal stress isassumed instead of σzz = 0 (Clasen et al. 2006a). Clasenet al. (2006a) stated that σzz increases with the sametime constant as D(t). Then Eq. 11 has to be corrected

Page 4: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

912 Rheol Acta (2012) 51:909–923

by a factor of 4−1/3 ≈ 0.63, but the elongational relax-ation time λe remains unaffected.

There is no simple universal relationship betweenthis characteristic elongational relaxation time andshear relaxation time. For a series of PIB solutions,λe ≈ 3λs was reported, with λs defined as an averageshear relaxation time (Liang and Mackley 1994). For PSBoger fluids with concentrations c ≈ c*, elongationalrelaxation times λe close to the Zimm relaxation timeλZ were found (Bazilevskii et al. 1997); but on theother hand, it was clearly shown for PEO as well asPS solutions that λe can vary drastically even at c < c*and λe/λZ values between 0.1 and 10 have been doc-umented (Tirtaatmadja et al. 2006; Christanti andWalker 2001a, b; Clasen et al. 2006a, b). Oliveira et al.(2006) related λe to the longest relaxation time λs

estimated from small amplitude oscillatory shear andfound λe ≈ λs PEO solutions with c ≈ c*. Arnolds et al.(2010) observed that λe/λs ≤ 1 and strongly decreaseswith increasing c for PEO solutions with c* < c < ce

and attributed this to the large deformation the solu-tions experience during filament thinning. They useda simple factorable integral model including a singlerelaxation time and a damping function to calculateλe/λs and obtained good agreement with experimentalresults.

Other more complex systems like surfactant solu-tions forming entangled wormlike micelles, polyelec-trolyte complexes in solution or aggregated acrylicthickener solutions also exhibit relaxation time ratiosλe/λs < 1, but these systems are supposed to undergostructural changes in strong elongational flow (Bhardwajet al. 2007; Kheirandish et al. 2008; Willenbacher et al.2008).

Finally, it should be mentioned that PEO solutionswith c > ce still form cylindrical filaments, but thetime evolution of the filament diameter is no longerexponential and cannot be characterized by a singlerelaxation time λe (Arnolds et al. 2010).

Determination of the axial force in a horizontallystretched filament

A fluid filament is bent when gravity is acting in radialdirection and the axial force within the liquid can becalculated from the bending line. Figure 2 shows asection of a liquid filament with infinitesimal length,ds, and the acting gravity force, dFG. The axial forceis generally not constant, which is considered in thecontribution of the infinitesimal force, dF, between theleft and right edge of the pictured thread. In order toevaluate the axial force, all forces are balanced with

Fig. 2 Forces in a tilted liquid thread (gray) with infinitesimallength ds

respect to a (x, y) coordinate system fixed in space.Figure 3 presents the constructional realization of thetilted CaBER method and the bent liquid filament.

It is clearly shown in Fig. 2 that the force balances inx- and y-direction results in dFx = 0 or Fx = C1, whereC1 is a constant of integration and dFy = dFG. Thegravity force is proportional to the volume dV and iscalculated assuming constant diameter D and density ρ:

dFG = g dm = ρg dV = π

4ρgD2 ds. (12)

Fig. 3 a The constructional realization of the tilted CaBERmethod. b The bent liquid filament

Page 5: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 913

Furthermore, ds can be written as

ds =√

(dx)2 + (dy)2 = dx

√1 +

(dydx

)2

(13)

and finally

dFy

dx= dFG

dx= π

4ρgD2

√1 + (y′)2 (14)

where y′ is the first derivative of the bending line y withrespect to x. The left side of Eq. 14 is obtained from thegeometrical ratio

Fx

Fy= dx

dy⇒ dFy

dx= Fx y′′. (15)

Combining Eqs. 14 and 15 yields the differentialequation for the bending line of the so called torque-free curved beam (see also Gross et al. 2009):

y′′ = πρgD2

4Fx

√1 + (y′)2. (16)

We use the substitutions y′ = u and y′′ = u′ to solvethe differential Eq. 16, and we choose the position ofmaximum deflection as point of reference to calculatethe constants of integration. Thus, we get the secondderivative of the bending line y′′ as a function of thex-position and the force component Fx:

y′′ = πρgD2

4Fxcosh

(πρgD2

4Fxx)

. (17)

The absolute value of the force F can be calculatedas a function of Fx and x analogous to the approachtaken in Eq. 13, using Eq. 15 and the identity cosh2(ϕ) −sinh2(ϕ) = 1.

F = Fx cosh(

πρgD2

4Fxx)

(18)

Equations 17 and 18 can be developed in Taylor series.Taking into account the first two terms (O(2)), the axialforce F is no longer a function of the position x and isgiven by

F = πρg4w′′ D2. (19)

For sake of clarity, we use the symbol w for theapproximated bending line and w′′ for its secondderivative.

The data evaluation is based on a least squareapproximation of the bending line by using a secondorder polynomial and calculating the constant secondderivative w′′ to be inserted in Eq. 19. However, a

constant diameter is assumed so far. This disadvantagecan be avoided using the deflection of the liquid threadinstead of the curvature of the bending line. Then, in-tegrating Eq. 19 twice yields the following relationshipbetween the axial force and the diameter D(t, x):

F = πρg4w

x∫0

x̃∫0

D(x̂)2dx̂dx̃. (20)

The double integral in Eq. 20 can be solved numeri-cally without any fitting of the bending line or any otherassumptions.

We estimate the approximation error, which may bedefined as follows:

β = 1 − w/y (21)

where y is the true and w the approximated bendingline. For given F and D, Fx is calculated numericallyin order to estimate the approximation error β. Linesof constant β are shown in Fig. 4 for a cylindricalfluid filament with density ρ = 1 g/cm3 and length2x = 12 mm. For the experiments presented here, β

was always below 0.1 %. It should be noted that thelength of the filament has a drastic influence on theapproximation error. For example, doubling the lengthof the filament to 24 mm while leaving the other para-meters unchanged results in 1 % < β < 10 %.

0 200 400 600 800 10000,01

0,1

1

10

100

1000

β = 1.0%

β = 0.5%

β = 0.2%

β = 0.1%

β = 0.1% β = 0.2% β = 0.5% β = 1.0%

forc

e F

[µN

]

diameter D [µm]

Fig. 4 Lines of constant approximation error β = 1 − w/y in theforce-diameter plane

Page 6: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

914 Rheol Acta (2012) 51:909–923

Experimental realization

The (tilted) CaBER device

The tilted CaBER method has been realized bymounting a commercial CaBER-1 (Thermo Scientific,Karlsruhe) on a customized L-shaped aluminum con-struction in order to rotate the whole device by 90◦, asshown in Fig. 3a. Thus, the liquid thread is stretchedin horizontal direction and bents due to the action ofgravity (Fig. 3b).

The device is equipped with plates of diameterD0 = 6 mm. The liquid bridge is created by separatingthe plates from a displacement of hi = 0.51 mm orhi = 0.75 mm to a final displacement of hf =6.5 mm orhf =11.5 mm within a constant strike time of ts =40 ms.

The upper plate or, in case of the tilted experi-ment, the left plate reaches its final position at timet1 = t0 + ts where the filament diameter is referred toD1 = D(t1). This notation considers the mismatch be-tween time t0 when the experiment starts and time t1when the capillary thinning begins. The thinning ofthe filament is recorded using a high-speed camera(Photron Fastcam-X 1024 PCI), a telecentric objec-tive (MaxxVision TC4M 16, magnification: ×1) and ablue telecentric background light (Vision & Control,TZB30-B) as described by Niedzwiedz et al. (2009).The experimental setup allows observations within aresolution of 1,024 × 1,024 px (1 px ≈ 16 μm) and aframe rate of 1,000 fps. We assume that at least twopixels are needed for a reliable determination of thefilament diameter, and we do not consider diameterslower than 32 μm. Each individual image is analyzedin order to determine the upper and lower edge ofthe filament. Here, we have taken care that edges aredetected properly correct without any impaired results,which can occur due to pixel fault (too bright or darkpixel) of the camera or defective image recognition.Therefore, we have implemented an automatic proce-dure in MATLAB to examine each image individu-ally relating to two aspects: the absolute value of thedetermined filament diameter must be lower than thediameter of the plates used and the maximum accept-able slope of the diameter is (dD/dx)max = 0.1. All datapoints which do not satisfy these conditions are ignoredin subsequent calculations.

Therewith, the experimental determined bendingline or neutral axis (see also Fig. 3b) is calculated toapply Eqs. 19 and 20 for the force calculation Theneutral axis is named W, this experimental value cor-responds to the model value w in the case of theapproximated theory and y in the case of the generaltheory. Regardless of the applicability of Eqs. 19 and

20, which will be discussed in more details below, theforce calculation from the bending line (Eq. 20) re-quires a further treatment of the measured data. Theinsufficient horizontal adjustment of the CaBER devicein a range of a few micrometers and a shear flowperpendicular to the axial filament direction (doublearrows in Fig. 3b) in the contact areas between thefluid filament and the CaBER-plates (which can differfor each plate) induced a different position of filamentfor the left and the right edge in the tilted experiment.Therefore, we carried out a baseline correction for ourdata, which is based on the assumption of filamentsymmetry relative to the point of maximum deflectionand contains the following strategy. We calculate thefirst derivative of the filament diameter dD/dx anddetermine the section where dD/dx < 0.01. The twodiscrete dD/dx values which are nearest to 0.01 arenamed L1 for the left side and R1 for the right sideof the filament (Fig. 5a). These values define the bor-ders for the baseline correction. Then the next fivepoints in positive x-direction are considered, and the

0

1

3 4 5 6 7 8 9 10 11 12 13 14 151

2

3

4

R1

dD/d

x [ %

]

L1

A

L1

bend

ing

line

W [m

m]

x position [mm]

L1+5 baseline

B

Fig. 5 a Derivation of the diameter with the determined leftedge, L1, and right edge, R1. b Experimentally determined bend-ing line W as a function of x-position. The gray lines representthe baseline BL calculated from L1 to L1+5 and R1−5 to R1 datapoints

Page 7: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 915

average value WL = ∑5i=0 W(L1+i)/6 is calculated; then

WR = ∑5i=0 W(R1−i)/6 is obtained analogously aver-

aging over six points starting from W(R1) going intonegative x-direction. Finally, the baseline BL is definedas the line connection of WL and WR (Fig. 5b). Thenthe corrected bending line Wcorr is calculated accordingto Wcorr = W − BL and the lowest point of Wcorr ischosen as point of reference so that (Wcorr)min = 0 isguaranteed. Finally, the force is calculated applyingEq. 20. It has to be noted that the baseline correctionhas no visible influence for the force calculation basedon Eq. 19.

Sample preparation and characterization

In this study, we used two different Newtonian modelsystems and one viscoelastic system. The first New-tonian system was composed of fructose (Carl RothGmbH, Karlsruhe, Germany) solutions with mass frac-tions between 76 and 81 % in an aqueous mixtureof 0.1 wt.% Tween20 (Carl Roth GmbH, Karlsruhe,Germany). The samples were stirred and heated for3 h using a magnetic stirrer and were measuredwithin 2 days to prevent recrystallization of the highlyconcentrated solutions (c ≥ 80 %). Zero shear vis-cosities η0 were determined from steady state mea-surements using a MARS II rotational rheometer(Thermo Scientific) equipped with cone plate geometry

(60/35 mm in diameter and 1◦ cone angle) at 20 ◦C.The surface tension (measured with a platinum–iridiumWilhelmy-plate mounted on a DCAT 11, DataPhysicstensiometer) was constant for all samples within exper-imental error and was determined to � ≈ 62.8 mN/m,which is approximately 20 mN/m lower than the surfacetension of the pure aqueous fructose solutions. There-fore, the addition of surfactant has two effects: thefilament lifetime of the fructose solutions is increasedand the surface tension exhibits the same value as forthe PEO solutions describe below. The second New-tonian system were mixtures of silicon oil AK50000in AK1000 (both Wacker Chemie AG, Munich,Germany) with mass fractions between 0 and 100 %.Silicon oils are obviously non-Newtonian fluids, butthey exhibit Newtonian flow behavior in CaBER ex-periments because the longest relaxation time λs ismuch smaller than the filament lifetime (compare Fig. 8and data in Table 1). The elasto-capillary number Ec(Anna and McKinley 2001; Clasen 2010) provides aquantitative criterion to estimate this effect.

Ec = 2λs�

ηs D(22)

It compares the viscous timescale tv = ηs D/2� withthe timescale of elastically controlled thinning, assumedas the longest shear relaxation time λs which was

Table 1 Physical properties of the measured model systems

Type c/% ηs,0/Pa s λs/ms �/mN/m ρ/g cm−3 Ecmax

Fructose/Tween20 76.0 0.88 ± 0.01 – 61.8 ± 2.3 1.380 –77.0 1.19 ± 0.03 – 64.4 ± 1.9 1.385 –78.0 1.64 ± 0.08 – 62.2 ± 0.7 1.390 –79.2 2.47 ± 0.05 – 61.9 ± 0.3 1.398 –80.0 3.59 ± 0.07 – 62.0 ± 0.8 1.402 –81.0 5.00 ± 0.11 – 62.4 ± 0.6 1.420 –

AK50000 0 1.04 ± 0.01 0.19 ± 0.05 20.8 ± 0.5 0.972 0.220 2.97 ± 0.01 0.82 ± 0.07 0.440 8.11 ± 0.07 1.71 ± 0.03 0.360 17.68 ± 0.23 3.26 ± 0.05 0.280 34.11 ± 0.41 4.63 ± 0.12 0.2

100 60.27 ± 1.18 5.97 ± 0.10 0.1PEO 106 g/mol 1.0 0.04 ± 0.01 9.7 ± 1.3 61.7 ± 0.6 0.992 935

1.5 0.14 ± 0.02 26 ± 2 62.7 ± 0.3 0.992 7282.0 0.38 ± 0.01 54 ± 3 61.7 ± 0.4 0.991 5482.5 0.94 ± 0.02 110 ± 8 62.8 ± 0.1 0.994 4593.0 2.01 ± 0.06 189 ± 8 62.8 ± 0.6 0.993 3693.5 4.10 ± 0.17 340 ± 15 62.5 ± 0.2 0.994 3244.0 8.58 ± 0.30 425 ± 21 62.8 ± 0.4 0.996 1944.5 12.97 ± 0.43 635 ± 27 61.0 ± 0.8 1.000 1875.0 25.46 ± 0.66 1005 ± 31 57.5 ± 0.5 1.005 142

Page 8: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

916 Rheol Acta (2012) 51:909–923

determined by small amplitude oscillatory shear exper-iments according to

λs = limω→0

G′

G′′ω(23)

using a Physica MCR 501 (Anton Paar) equipped withcone plate geometry (50-mm diameter and 1◦ coneangle) at 20 ◦C.

Elastically driven capillary thinning sets in whenEc ≈ 1. For the silicon oil mixtures investigated here,the maximum elasto-capillary number ECmax werecalculated using the minimal observable diameter ofDmin = 32 μm. In all cases investigated here, Ecmax 1was detected, corresponding data are summarized inTable 1.

Therefore, silicon oils are a useful Newtonian modelsystem for CaBER experiments. The surface tensionwas only measured for pure AK1000 due to the highviscosity of the other samples, and we assumed a con-stant surface tension for all silicon oil mixtures.

PEO solutions with polymer mass fractions between1 and 5 % were used as viscoelastic model system.The polymer powder (Aldrich Chemical Co., UK) hada weight average molecular weight of 106 g/mol andwas dissolved in distilled water by means of shaking.The samples were stored at 7 ◦C up to the point ofmeasurement. The surface tension of the PEO solutionsis � ≈ 62.4 mN/m and independent of polymer concen-tration up to c ≈ 4 %. A lower surface tension wasobserved for higher polymer concentration, probablydue to systematic errors inferred from the high sampleviscosity.

All data, including the density of the samples de-termined using a pycnometer with a total volume of10.706 cm3 at 20 ◦C, are summarized in Table 1.

Results and discussion

Validation of the force calculation method

Figure 6 shows typical results of force calculation fromgravity-driven bending of horizontally stretched fluidfilaments as a function of diameter. The squares rep-resent the 2.5 % PEO solution which behaves like atypical weakly viscoelastic fluid, indicated by a cylin-drical thread in the CaBER experiment. The forces cal-culated using Eq. 20 (open symbols) and Eq. 19 (closedsymbols) are identical within experimental uncertainty.Therefore, the simplified force calculation method, andparticularly the parabolic fit of the bending line, yieldsgood results as long as the thread is cylindrical.

250 200 150 100 50 00.1

1

10

100

Eq. 20 Eq. 19 PEO 2.5% AK50000

forc

e F

[µN

]diameter D [µm]

Fig. 6 Measured force F as a function of the diameter for the endof the filament thinning (measurement properties, hi = 0.51 mm,hf = 11.5 mm). Closed symbols represent data evaluation, usingEq. 20; and open symbols represent data evaluation, using Eq. 19,for a 2.5 % PEO solution (squares) and silicon oil AK50000(triangles)

The Newtonian silicon oil AK50000 forms slightlyconcave filaments. The calculated forces (triangles inFig. 6) differ from each other due to the non-constantdiameter. The force values calculated using the numeri-cal integration of the diameter (Eq. 20), represented byclosed symbols, are significantly larger than the forcevalues calculated by Eq. 19 with the assumption ofconstant diameter (open symbols). The correspondingTrouton ratios Tr defined as the ratio of the true elon-gational viscosity, calculated by applying Eq. 7, and theshear viscosity are Tr = 3.2 when the force is calculatedaccording to Eq. 20 and Tr = 2.0 if Eq. 19 is applied.Of course, the correct Trouton ratio is three for a New-tonian fluid. This demonstrates that the filament cur-vature has to be taken into account for non-cylindricalfilaments and as a consequence, we determine the forceonly by applying Eq. 20 for all cylindrical and non-cylindrical threads.

Figure 7 (left) shows the force ratio X as a functionof diameter calculated directly from the force mea-surements according to Eq. F2 for three silicon oilmixtures. For all samples investigated here, X stronglydecreases with decreasing diameter to a constant finalvalue X∞. This decrease is associated with the initialdeflection of the liquid tread. The critical filamentdiameter, where X∞ is reached, is shifted to lower

Page 9: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 917

Fig. 7 Left Force ratio as afunction of diameter forAK50000 in AK1000 mixturefor three differentconcentrations given in thediagram. Right Imagesequence of the deflectionduring the capillary breakupin the tilted CaBERexperiment for the 20 %AK50000 mixture

500 400 300 200 100 00

1

2

C

AK 50000 20% 60% 80%

forc

e ra

tio X

[ -

]

diameter D [µm]

X∞

A

BD

E

values with increasing AK50000 concentration and cor-respond to the inflection point of the filament diametervs. time curves, shown in Fig. 8. The low viscous siliconoil mixture including of 20 % AK50000 additionallyexhibits a sharp minimum before the constant X∞ valueis reached. This initial variation of X is attributed toa wave traveling through the filament after the plateshave reached their final position. This wave results in ashift of the point of maximum deflection as shown in theimages in Fig. 7. Image D where the filament deflectionis symmetric corresponds to the diameter where X∞ isreached. This finding is typical for low viscous New-tonian fluids with ηs < 15 Pa s. For the viscoelastic

0 10

100

200

300

400

500

0 1 2 3 0 1 2 3 4 5

diam

eter

D [µ

m]

60% 80% 100%

time t [s]

s,0

Γ−Θη

0% 20% 40%

Fig. 8 Diameter as a function of time for the AK50000 inAK1000 mixtures (hi = 0.51 mm, hf = 11.5 mm)

PEO solutions investigated here, a wave through thefilament was only observed at a zero shear viscosityof ηs = 0.04 Pa s, corresponding to a concentration of1.0 %. This phenomenon did not occur for solutionswith higher PEO concentrations and this is attributedto the well-known stabilization effect of the stretchedpolymer coils (Lenczyk and Kiser 1971; Tirtaatmadjaet al. 2006).

The consideration of the final force ratio allows foranother way for the validation of the force calculationmethod. On one hand, X∞ is determined directly fromthe force measurement as described above. On theother hand, X∞ can also be calculated from the slopeof D(t) which is proportional to −��/ηs (see Fig. 8),as expected for Newtonian fluids. Using Eq. 10 and theadditionally measured zero shear viscosity, the numer-ical factor � is linked to the force ratio X∞. Figure 9shows the experimentally determined force ratios, as afunction of viscosity for the two series of Newtonianfluids. Here, silicon oil mixtures (squares) were mea-sured with an initial height of hi = 0.51 mm and afinal height of hf = 11.5 mm and the fructose/Tween20solutions (triangles) were measured with hi = 0.75 mmand hf = 6.5 mm. The large scale error bars of thefructose/Tween20 solutions are caused by the highcapillary velocity �/ηs, and, hence, a short filament life-time. The force ratio values determined from the diam-eter (open symbol) and the force (filled symbols) usingEq. 20 are identical within experimental error. Con-sidering all force measurement experiments, we obtainan average force ratio X∞ = 0.728 ± 0.059, which is in

Page 10: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

918 Rheol Acta (2012) 51:909–923

1 10 1000.5

0.6

0.7

0.8

0.9

1.0

solution of Eggers/Brenner

from F from D silicon oil fructose/Tween20

final

forc

e ra

tio X

[ - ]

zero shear viscosity [Pas]

solution of Papageorgiou

Fig. 9 Force ratio as a function of zero shear viscosity for siliconoil mixtures (hi = 0.51 mm, hf = 11.5 mm) and fructose/Tween20solutions (hi = 0.75 mm, hf = 6.5 mm). The f illed symbols repre-sent X∞ values using the calculated force, and the open symbolsrepresent X∞ values using the diameter

excellent agreement with the solution of Papageorgiou(1995) and which is far away from the solution ofEggers (1993, 1997) and Brenner et al. (1996). Allthese experiments validate the tilted CaBER methodincluding the force calculation from Eq. 20 as anaccurate and robust method for the determination ofthe axial force in a fluid filament during capillary thin-ning elongational flow.

Sign of the axial normal stress

From a theoretical point of view, the minimum σzz/|σrr|value is −1 because lower values would result in neg-ative elongational viscosities. Figure 10 shows the di-mensionless normal stress ratio σzz/|σrr| as a functionof diameter for a silicon oil mixture of 60 % AK50000in AK1000 (filled squares) and a 3.5 % PEO solution(open triangles).

During the initial thinning period dominatedσzz/|σrr| strongly decays in both cases. This decay isfaster for fluids of lower viscosity and is presumably dueto inertial effects. After this induction period σzz/|σrr|reaches a constant value σzz/|σrr| = −0.54 ± 0.12 for allNewtonian fluids. The negative sign corresponds to acompressive stress, which causes the pronounceddeflections of Newtonian filaments. Such compressivestresses are not observed for the PEO solutions and,hence, the deflection is less pronounced comparedto the Newtonian samples discussed in the previoussection. The induction period is much shorter for all theinvestigated PEO solutions than for the silicon oilsdue to the damping properties of the polymer coils.

700 600 500 400 300 200 100 0-2

-1

0

1

2

AK50000 60% PEO 3.5%

norm

al s

tres

s ra

tio σ

zz/|σ

rr| [

- ]

diameter D [µm]

Fig. 10 Normal stress ratio for a 3.5 % PEO 106 g/mol solution(f illed squares) and a 60 % AK50000 mixture (open triangles)

Accordingly, the asymptotic value for the normalstress ratio is reached much faster. A limiting valueσzz/|σrr|≈0 is observed for the PEO solutions withconcentrations lower than 4 %. For these sparselyconcentrated solutions, it is not possible to check theassumption of Clasen et al. (2006a) that σzz increasesexponentially due to the measuring accuracy andthe scatter of the data points. Anyway, σzz must besignificantly smaller than σrr, and in this case, the widelyused assumption σzz = 0 obviously is valid for theevaluation of the elongational viscosity from CaBERmeasurements. But constant values σzz/|σrr| > 0are observed for higher polymer concentrationsas discussed in the next section. This finding isthe experimental proof of significant exponentiallyincreasing axial normal stresses as predicted by Clasenet al. (2006a).

σzz

|σrr| = const. ⇒ σzz ∝ 1D

∝ exp(

t3λe

)(24)

The normal stress ratio is related to the force ratio Xdiscussed in the previous section. The force balance(Eqs. 1 and 3) yields:

σzz

|σrr| = 2(

Fπ D�

− 1)

. (25)

Page 11: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 919

Combining Eq. 25 with the definition of the forceratio (Eq. 2) results in

X = 12

σzz

|σrr| + 1. (26)

Therefore, the sign of σzz is obvious from the forceratio, as discussed above. For vanishing axial stresses,the force ratio is one. X < 1 values correspond toa compressive axial normal stress, and X > 1 valuescorrespond to a tensile normal stress.

Applying the tilted CaBER methodto non-Newtonian fluids

Recently, the behavior of semi-dilute and concentratedentangled PEO solutions in regular CaBER experi-ments has been discussed in detail (Arnolds et al. 2010).They obtained c* ≈ 0.4 % and ce ≈ 2.5 % for PEOwith Mw = 106 g/mol from their steady and small am-plitude oscillatory shear experiments. Here, we focuson the variation of the axial force in this concentrationrange investigating PEO solutions with concentrationsbetween 1 and 5 %.

The filament lifetime of these PEO solutionsstrongly increases with polymer concentration. Diam-eter data are shown in Fig. 11a. The PEO solutionswith c < 4 % exhibit the typical exponential thinning(lines in Fig. 11a), with characteristic relaxation timeλe, calculated according to Eq. F11. For higher polymerconcentrations, the filaments are still cylindrical but donot exhibit an exponential thinning regime and, hence,λe cannot be determined. The findings are in goodagreement with Arnolds et al. (2010). Correspondingforce measurement data are shown in Fig. 11b. Theforce F decreases monotonically with time to a minimalvalue of about 1 μN.

Calculated force ratios as a function of diameter Dfor PEO solutions with c = 3.5 % and c = 5.0 % as wellas for the Newtonian silicon oil mixture AK50000 60 %are shown in Fig. 12a. The measured force ratios forthe PEO solutions investigated here are always higherthan the force ratio for the Newtonian liquid. Remark-ably, this finding disagrees with the observation ofClasen (2010) for capillary breakup of semi-dilute poly-mer solutions. He observed that the thinning processsplits up into four regimes: (1) early thinning whichis controlled by gravitational sagging; (2) Newtonianthinning behavior (viscocapillary) described by Eq. 9;(3) extensional thinning (still viscocapillary) describedby a power law; and finally (4) elasto-capillary thinningdescribed by Eq. 11. It is clearly shown in Fig. 12athat the X(D) data do not support the existence ofNewtonian thinning regime because the Newtonian

limit of X∞ = 0.728 ± 0.059 is never reached duringthe thinning process. This emphasizes that the inter-pretation of the D(t) data (Fig. 11a) as, e.g., done byClasen (2010) is not unambiguous since the correspond-ing force ratio is not considered. Therefore, the forcemeasurement gives additional information for the inter-pretation of the thinning behavior of liquid filaments.

0 1 210

100

1000 PEO 1.5% PEO 2.5% PEO 3.5% PEO 4.5% PEO 5.0%

diam

eter

D [µ

m]

time t [s]

A

0 1 21

10

100

1000 PEO 1.5% PEO 2.5% PEO 3.5% PEO 4.5% PEO 5.0%

forc

e F

[µN

]

time t [s]

B

Fig. 11 PEO solutions CaBER measurements (measurementproperties, hi = 0.51 mm, hf = 11.5 mm). a Diameter as a func-tion of time. The lines represent exponential fits to the data.b Force as a function of time

Page 12: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

920 Rheol Acta (2012) 51:909–923

The extracted final force ratios X∞ versus concen-tration c for the investigated PEO solutions are shownin Fig. 12b. Remarkably, this ratio is one for theclearly elasto-capillary exponentially thinning solutions

1000 800 600 400 2000

1

2

3

4

5

AK50000 60% PEO 3.5% PEO 5.0%

forc

e ra

tio X

[ -

]

diameter D [µm]

A

1 2 3 4 50

1

2

hi = 0.51mm

hi = 0.75mm

final

forc

e ra

tio X

∞ [

- ]

concentration c [%]

B

Fig. 12 a Force ratio as a function of diameter. b Final force ratioas a function of concentration c for two different initial heightshi = 0.51 mm (f illed squares) and hi = 0.75 mm (open triangles).The dashed line represents a vanishing axial normal stress

(c < 3.5 %). X∞ = 1 means that there is no axial stressin the filament during capillary thinning. Therefore,the simplified approach of calculating the elongationalviscosity according to Eq. 8 is valid for these weaklyelastic PEO solutions. However, the final force ra-tio deviates significantly from X∞ = 1, indicating thatσzz = 0 and increases monotonically for concentrationsc ≥ 4 %. Non-vanishing σzz values for non-Newtonianfluids are directly detected here for the first time per-forming CaBER experiments. This clearly indicates thelimitations of the widely used, simplified data analysisfor calculating the elongational viscosity (Eq. 8) anddemonstrates that an accurate force measurement ismandatory for a determination of absolute ηe values(Eq. 7) from CaBER experiments.

The concentration where X∞ begins to differ fromunity (c = 4 %) is much greater than the overlap con-centration of c* and roughly coincides with the entan-glement concentration ce. It seems that a significantaxial normal stress is due to the entanglements presentin these solutions and gives rise to the observed non-exponential thinning of the filament diameter.

In Fig. 13, the elongational viscosity ηe obtainedfrom the force measurement is compared to the ap-parent elongational viscosity ηe,app calculated from �

and D(t) using the σzz = 0 assumption for a 4.0 % PEOsolution (Fig. 13a) and a 5.0 % PEO solution (Fig. 13b).Here, we have smoothed the diameter data applyingthe floating average method in order to calculate theelongation rate numerically.

In the initial period of capillary thinning right afterthe upper plate has reached its final position (t > t1),the elongational viscosity ηe obtained from the forcemeasurements goes through a sharp maximum, whichdoes not appear in ηe,app. Similar results were ob-tained for all PEO solutions investigated here, but theapparent viscosity maximum is less pronounced forPEO solutions with higher polymer concentration.

The physical interpretation of this phenomenon isnot yet clear, but it might be due to inertial effects asalready discussed above (see “Validation of the forcecalculation method” and “Sign of the axial normalstress”). It also has to be noted that the elongationrate goes through a distinct minimum during this periodof thinning. This topic requires further investigationswhich are beyond the scope of this paper.

The time when ηe reaches its minimum correspondsto the time when the force ratio X reaches its limit-ing value. For the PEO solutions with c ≤ 4 %, thetime also corresponds to the inset of the exponentialthinning. In this case, the limiting value is independentof polymer concentration (X∞ = 1.1 ± 0.1). During thefinal stage of capillary thinning, both elongational vis-

Page 13: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 921

0.0 0.2 0.4 0.6 0.8 1.010

100

1000

4% PEO solution from force F from surface tension

elon

gatio

nal v

isco

sity

e [P

as]

time t [s]

t1

3s

A

0.0 0.5 1.0 1.5 2.010

100

1000

elon

gatio

nal v

isco

sity

e [P

as]

time t [s]

t1

3s

5% PEO solution from force F from surface tension B

ηη

η

η

Fig. 13 Elongational viscosity as a function of time (measure-ment parameters, hi = 0.75 mm, hf = 11.5 mm), calculated fromforce measurement data according to Eq. 7 (f illed squares) andthe surface tension (Eq. 8) assuming σzz = 0 (open triangles).a 4.0 % PEO solution. b 5.0 % PEO solution

cosity values ηe and ηe,app strongly increase during thecapillary thinning due to the well-known progressiveentropic resistance to the deformation of the polymerchains; but in this case, ηe,app is a reasonable approx-imation for the true elongational viscosity. For the5 % PEO solution with X∞ > 1, the apparent elon-gational viscosity during the final stage of thinning is

significantly lower than the true value obtained fromthe force measurements.

Conclusions

We have introduced a new way of performing CaBERexperiments named tilted CaBER method. Thismethod comprises a horizontal stretching of fluid fila-ments and allows for a determination of the axial forcein the liquid bridge from the gravity-driven bendingof the filament applying chain bending theory. Thismethod provides reliable values for the axial force ina liquid filament in a range of N = 1,000–0.1 μN andallows for a determination of the true elongational vis-cosity without additional assumptions or specific con-stitutive equations. For Newtonian fluids, the existenceof a negative axial normal stress was proved experi-mentally, and the solution of Papageorgiou (1995) wasconfirmed for the linear thinning of Newtonian liquidsby using two different sets of samples covering a viscos-ity range of ηs = 0.9–60 Pa s. The tilted CaBER methodhas been applied to non-Newtonian PEO solutions withconcentrations c > c* and c > ce. The assumption ofvanishing axial normal stress for evaluating CaBERexperiments was confirmed for solutions with c < ce,and a positive axial normal stress was detected forconcentrations above the entanglement concentration.In this case, the time evolution of the filament diameterno longer follows a simple exponential decay law. Theelongational viscosity calculated from the measuredforce exhibits the well-known elongational hardeningbehavior also observed for other polymer solutions(Sridhar et al. 1991; Solomon and Muller 1996).

Acknowledgements The authors would like to thank S. Bindgenfor his help in the sample preparation and performing CaBERexperiments. Financial support by German Research FoundationDFG grant WI 3138/13–1 is gratefully acknowledged.

Appendix

Equation 7 for the calculation of the elongationalviscosity in axial direction is only valid for slenderfilaments as investigated in this paper. Filament curva-ture must be taken into account in case of strongly non-cylindrical filaments which are typical for, e.g., yieldstress fluids. At the position of the neck, the equilib-rium condition in radial direction then reads as follows:

σrr = −�

(2

Dmid− 1

|Rc|)

(27)

Page 14: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

922 Rheol Acta (2012) 51:909–923

where Dmid is the diameter and Rc is the radius ofcurvature in axial direction at the neck. The expressionfor σzz does not change, and accordingly, the true elon-gational viscosity is given by

ηe = σzz − σrr

ε̇= 1

dDmid/dt

(� − 2F

π Dmid+ �Dmid

2 |Rc|)

.

(28)

References

Agarwal S, Gupta RK (2002) An innovative extensional viscome-ter for low-viscosity and low-elasticity liquids. Rheol Acta41:456–460

Alexandrou AN, Entov VM, Bazilevsky AV, Rozhkov AN,Sharaf A (2009) On the tensile testing of viscoplastic fluids.In: Acta rheological, special issue following the workshop:viscoplastic fluids: from theory to application, Nov. 1–5

Anna SL, McKinley GH (2001) Elasto-capillary thinning andbreakup of model elastic liquids. J Rheol 45:115–138

Arnolds O, Buggisch H, Sachsenheimer D, Willenbacher N(2010) Capillary breakup extensional rheometry (CaBER)on semi-dilute and concentrated polyethylene oxide (PEO)solutions. Rheol Acta 49:1207–1217

Bazilevskii AV, Entov VM et al (1997) Failure of polymer solu-tion filaments. Polym Sci A 39:316–324

Bazilevsky AV, Entov VM, Rozhkov AN (1990) Liquid filamentmicrorheometer and some of its applications. In: OliverDR (ed) Third European rheology conference. Elsevier,San Diego, pp 41–43

Bazilevsky AV, Entov VM, Rozhkov AN (2001) Breakup of anOldroyd liquid bridge as a method for testing the rheologicalproperties of polymer solutions. Polym Sci Ser A 43:716–726

Bhardwaj A, Miller E, Rothstein JP (2007) Filament stretchingand capillary breakup extensional rheometry measurementsof viscoelastic wormlike micelle solutions. J Rheol 51:693–719

Braun DB, Rosen MR (2000) Rheology modifiers handbook—practical use & application. William Andrew Publishing,Norwich, US

Brenner M, Lister J, Stone H (1996) Pinching threads, singulari-ties and the number 0.0304. Phys Fluids 8:2827–2836

Chao KK, Child CA, Grens EA, Williams MC (1984) Antimistingaction of polymeric additives in jet fuels. AIChE J 30:111–120

Christanti Y, Walker LM (2001a) Surface tension driven jetbreak up on strain-hardening solutions. J Non-Newton FluidMech 100:9–26

Christanti Y, Walker LM (2001b) Effect of fluid relaxation timeof dilute polymer solutions on jet breakup due to a forceddisturbance. J Rheol 46:733–748

Clasen C (2010) Capillary breakup extensional rheometry ofsemi-dilute polymer solutions. K-A Rheol J 22:331–338

Clasen C, Eggers J, Fontelos MA, Li J, McKinley GH (2006a)The beads on-string structure of viscoelastic threads. J FluidMech 556:283–308

Clasen C, Plog JP et al (2006b) How dilute are dilute solutions inextensional flows? J Rheol 50:849–881

Dexter RW (1996) Measurement of extensional viscosity of poly-mer solutions and its effects on atomization from a spraynozzle. At Sprays 6:167–191

Eggers J (1993) Universal pinching of 3D axisymmetric free-surface flows. Phys Rev Lett 71:3458–3490

Eggers J (1997) Nonlinear dynamics and breakup of free-surfaceflows. Rev Mod Phys 69:865–929

Entov VM, Hinch EJ (1997) Effect of a spectrum of relaxationtimes on the capillary thinning of a filament of elastic liquid.J Non-Newton Fluid Mech 72:31–53

Fernando RH, Lundberg DJ, Glass JE (1989) Importance ofelongational flows in the performance of water-borne for-mulations. Adv Chem 223:245–259

Fernando RH, Xing LL, Glass JE (2000) Rheology parameterscontrolling spray atomization and roll misting behavior ofwaterborne coatings. Prog Org Coat 40:35–38

Gross D, Hauger W, Wriggers P (2009) Technische mechanikband 4: hydromechanik, elemente der höheren mechanik.Numerische Methoden Springer, Berlin

Han L, Gupta RK, Doraiswamy D, Palmer M (2004) Effect ofliquid rheology on jetting of polymer solutions. In: XIVthint congr rheol, Seoul, KR

James DF, Yogachandran N, Roper JA III (2003) Fluid elasticityin extension, measured by a new technique, correlates withmisting. In: 8th TAPPI advanced coating fundamentals sym-posium, Chicago, US, pp 166–171

Kennedy JC, Meadows J, Williams PA (1995) Shear andextensional viscosity characteristics of a series of hydropho-bically associating polyelectrolytes. J Chem Soc FaradayTrans 91:911–916

Kheirandish S, Gubaydullin I, Willenbacher N (2008) Shear andelongational flow behaviour of acrylic thickener solutions,Part I: effect of intermolecular aggregation. Rheol Acta49:397–407

Klein CO, Naue IF et al (2009) Addition of the force mea-surement capability to a commercially available extensionalrheometer (CaBER). Soft Mater 7:242–257

Kolte MI, Szabo P (1999) Capillary thinning of polymericfilaments. J Rheol 43:609–626

Lenczyk JP, Kiser KM (1971) Stability of vertical jets of non-Newtonian fluids. AIChE J 17:826–831

Liang RF, Mackley MR (1994) Rheological characterization ofthe time and strain dependence for polyisobutylene solu-tions. J Non-Newton Fluid Mech 52:387–405

McKay GR, Ferguson J, Hudson NE (1978) Elongational flowand the wet spinning process. J Non-Newton Fluid Mech4:89–98

McKinley GH, Tripathi A (2000) How to extract the Newtonianviscosity from capillary breakup measurements in a filamentrheometer. J Rheol 44(3):653–670

Meadows J, Williams PA, Kennedy JC (1995) Comparisonof extensional and shear viscosity characteristics of aqueoushydroxyethyl cellulose solutions. Macromol 28:2683–2692

Ng SL, Mun RP, Boger DV, James DF (1996) Extensional vis-cosity measurements of dilute solutions of various polymers.J Non-Newton Fluid Mech 65:291–298

Niedzwiedz K, Arnolds O, Willenbacher N, Brummer R (2009)How to characterize yield stress fluids with capillary breakupextensional rheometry (CaBER)? Appl Rheol 19:41969–1–41969–10

Oliveira MSN, Yeh R, McKinley GH (2006) Iterated stretch-ing, extensional rheology and formation of beads-on-stringstructures in polymer solutions. J Non-Newton Fluid Mech137:137–148

Papageorgiou DT (1995) On the breakup of viscous liquidsthreads. Phys Fluids 7:1529–1544

Plog JP, Kulicke WM, Clasen C (2005) Influence of the molarmass distribution on the elongational behaviour of polymersolutions in capillary breakup. Appl Rheol 15:28–37

Page 15: Determination of axial forces during the capillary breakup of liquid … · 2012. 10. 26. · Rheol Acta (2012) 51:909–923 911 It is obvious from Eq. 7 that for calculating η e,the

Rheol Acta (2012) 51:909–923 923

Prud’homme RK, Smith-Romanogli V, Dexter RW (2005) Elon-gational viscosity effects in spraying processes. In: 230thACS national meeting, Washington, US

Schümmer P, Tebel KH (1983) A new elongational rheometerfor polymer solutions. J Non-Newton Fluid Mech 12(3):331–347

Solomon MJ, Muller SJ (1996) The transient extensionalbehavior of polystyrene-based Boger fluids of varyingsolvent quality and molecular weight. J Rheol 40:837–856

Sridhar T, Tirtaatmadja V, Nguyen DA, Gupta RK (1991)Measurement of extensional viscosity of polymer solutions.J Non-Newton Fluid Mech 40(3):271–280

Stelter M, Brenn G, Yarin AL, Singh RP, Durst F (2002)Investigation of the elongational behavior of polymer so-lutions by means of an elongational rheometer. J Rheol46:507–527

Tan H, Tam KC, Tirtaatmadja V, Jenkins RD, Bassett DR (2000)Extensional properties of model hydrophobically modifiedalkali-soluble associative (HASE) polymer solutions. J Non-Newton Fluid Mech 92:167–185

Tirtaatmadja V, McKinley GH, Cooper-White JJ (2006) Dropformation and breakup of low viscosity elastic fluids:effects of molecular weight and concentration. Phys Fluids18:043101

Vadillo DC et al (2010) Evaluation of ink jet fluid’s perfor-mance using the “Cambridge Trimaster” filament stretchand breakup-device. J Rheol 54(2):261–282

Webster MF, Matallah et al (2008) Numerical modeling of Stepp-strain for stretched filaments. J Non-Newton Fluid Mech151:38–58

Willenbacher N, Matter Y, Gubaydullin I (2008) Effect of aggre-gation on shear and elongational flow properties of acrylicthickeners. K-A Rheol J 20:109–116