12
Influence of patterned surface in the rheometry of simple and complex fluids q Diana Broboana, Nicoleta Octavia Tanase, Corneliu Balan ‘‘Politehnica’’ University of Bucharest, REOROM Laboratory, Hydraulics Department, Splaiul Independentei 313, 060042 Bucharest, Romania article info Article history: Received 29 May 2014 Received in revised form 8 October 2014 Accepted 15 October 2014 Available online 24 October 2014 Keywords: Rheometry Patterned surfaces Microgeometries Apparent slip Numerical simulations abstract The influence of the surface microgeometry in rheometry is investigated using the correlation between experiments and numerical simulations of the flow in plate–plate rotational geometry. The presence of micropillars or microchannels on the plates induces effective or apparent slip of the fluids at the walls generating a ‘‘dynamic hydrophobic surface’’. The end/edge effects and the gap error are first estimated for the plate–plate smooth geometry, the computational results being found consistent with the per- formed experiments. The comparison between the measured torques in smooth and microchannels pat- terned configurations are analyzed using the numerical simulations performed by the Fluent code for the Newtonian fluid and the Carreau model. The present study demonstrates that hydrophobic effects can be induced at the walls, without the vio- lation of the no-slip condition, by changing the local flow spectrum due to the presence of patterned sur- faces at the solid boundary. The results confirm that computational rheometry is an useful tool not just to interpret the experimental data but to calculate the errors of the measurements, as well as to explore and model flow phenomenon as the apparent slip. The applications of the paper are meant to develop novel testing procedures in rheometry and to design micro-patterned surfaces for the control of slip/adherence of simple and complex liquids in microfluidic devices. Ó 2014 Elsevier B.V. All rights reserved. 1. Introduction Microfluidics is today one of the most dynamic domain of study in engineering, especially in relation with novel Lab-on-a-Chip applications which involve the flows of simple and complex fluids [1–3]. One important subject of investigation in microfluidics is the prediction of the fluid behavior in the very vicinity of the walls. Is the fluid slipping or not to the wall? [4,5]; is the surface hydro- philic or hydrophobic in respect to a particular dynamic process? [6–8]; which is the most indicated pattern of the wall to induce hydrophobicity? [9–11], are inevitable questions whose answers would be an important impact in the design procedure of the new devices and applications [12]. Therefore, the control of wall adherence and degree of slipping become central topics for the fundamental studies in microfluidics. If the material surface and the fluid sample are well defined, in order to control the degree of fluid adherence to the walls then we have to design the proper micropattern which induces the desired phenomenon (i.e. total/ partial adherence or slipping). The rheological bulk properties of liquids are experimentally determined using rheometers and spe- cial designed measurements procedures based on viscometric flows [13]. In all commercial rheometers (rotational or capillary), the solid surfaces in contact with tested liquids are normally smooth and the slip or perfect adherence of the samples to these surfaces are analyzed and interpreted for each type of measure- ment, mainly in relation with the microstructure and formulation of the liquid samples [14–17]. The goal of the present study is to investigate and analyze the influence of patterned surfaces in rheometry, to model the flow in plate and plate rotational geometry and to understand how the presence of microgeometries on the plate’s surface induces the apparent slipping at the wall and creates ‘‘dynamic hydropho- bic surfaces’’, even if the fluid is considered to adhere to the solid walls. In 1975, professor Ken Walters published Rheometry [18], the first book dedicated exclusively to the measurements procedures of the rheological properties of non-Newtonian fluids. In Introduc- tion, the author presented one main objective of this discipline (which is working in ‘‘tandem’’ with rheological modeling and numerical simulations [19,20]): ‘‘to determine the behavior of http://dx.doi.org/10.1016/j.jnnfm.2014.10.006 0377-0257/Ó 2014 Elsevier B.V. All rights reserved. q Dedicated to Prof Ken Walters FRS on the occasion of his 80th Birthday. Corresponding author. Tel./fax: +40 21 4029865. E-mail address: [email protected] (C. Balan). Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162 Contents lists available at ScienceDirect Journal of Non-Newtonian Fluid Mechanics journal homepage: http://www.elsevier.com/locate/jnnfm

Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

Contents lists available at ScienceDirect

Journal of Non-Newtonian Fluid Mechanics

journal homepage: ht tp : / /www.elsevier .com/locate / jnnfm

Influence of patterned surface in the rheometry of simple and complexfluids q

http://dx.doi.org/10.1016/j.jnnfm.2014.10.0060377-0257/� 2014 Elsevier B.V. All rights reserved.

q Dedicated to Prof Ken Walters FRS on the occasion of his 80th Birthday.⇑ Corresponding author. Tel./fax: +40 21 4029865.

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

Diana Broboana, Nicoleta Octavia Tanase, Corneliu Balan ⇑‘‘Politehnica’’ University of Bucharest, REOROM Laboratory, Hydraulics Department, Splaiul Independentei 313, 060042 Bucharest, Romania

a r t i c l e i n f o

Article history:Received 29 May 2014Received in revised form 8 October 2014Accepted 15 October 2014Available online 24 October 2014

Keywords:RheometryPatterned surfacesMicrogeometriesApparent slipNumerical simulations

a b s t r a c t

The influence of the surface microgeometry in rheometry is investigated using the correlation betweenexperiments and numerical simulations of the flow in plate–plate rotational geometry. The presence ofmicropillars or microchannels on the plates induces effective or apparent slip of the fluids at the wallsgenerating a ‘‘dynamic hydrophobic surface’’. The end/edge effects and the gap error are first estimatedfor the plate–plate smooth geometry, the computational results being found consistent with the per-formed experiments. The comparison between the measured torques in smooth and microchannels pat-terned configurations are analyzed using the numerical simulations performed by the Fluent code for theNewtonian fluid and the Carreau model.

The present study demonstrates that hydrophobic effects can be induced at the walls, without the vio-lation of the no-slip condition, by changing the local flow spectrum due to the presence of patterned sur-faces at the solid boundary. The results confirm that computational rheometry is an useful tool not just tointerpret the experimental data but to calculate the errors of the measurements, as well as to explore andmodel flow phenomenon as the apparent slip.

The applications of the paper are meant to develop novel testing procedures in rheometry and to designmicro-patterned surfaces for the control of slip/adherence of simple and complex liquids in microfluidicdevices.

� 2014 Elsevier B.V. All rights reserved.

1. Introduction

Microfluidics is today one of the most dynamic domain of studyin engineering, especially in relation with novel Lab-on-a-Chipapplications which involve the flows of simple and complex fluids[1–3]. One important subject of investigation in microfluidics is theprediction of the fluid behavior in the very vicinity of the walls. Isthe fluid slipping or not to the wall? [4,5]; is the surface hydro-philic or hydrophobic in respect to a particular dynamic process?[6–8]; which is the most indicated pattern of the wall to inducehydrophobicity? [9–11], are inevitable questions whose answerswould be an important impact in the design procedure of thenew devices and applications [12]. Therefore, the control of walladherence and degree of slipping become central topics for thefundamental studies in microfluidics. If the material surface andthe fluid sample are well defined, in order to control the degreeof fluid adherence to the walls then we have to design the propermicropattern which induces the desired phenomenon (i.e. total/

partial adherence or slipping). The rheological bulk properties ofliquids are experimentally determined using rheometers and spe-cial designed measurements procedures based on viscometricflows [13]. In all commercial rheometers (rotational or capillary),the solid surfaces in contact with tested liquids are normallysmooth and the slip or perfect adherence of the samples to thesesurfaces are analyzed and interpreted for each type of measure-ment, mainly in relation with the microstructure and formulationof the liquid samples [14–17].

The goal of the present study is to investigate and analyze theinfluence of patterned surfaces in rheometry, to model the flowin plate and plate rotational geometry and to understand howthe presence of microgeometries on the plate’s surface inducesthe apparent slipping at the wall and creates ‘‘dynamic hydropho-bic surfaces’’, even if the fluid is considered to adhere to the solidwalls.

In 1975, professor Ken Walters published Rheometry [18], thefirst book dedicated exclusively to the measurements proceduresof the rheological properties of non-Newtonian fluids. In Introduc-tion, the author presented one main objective of this discipline(which is working in ‘‘tandem’’ with rheological modeling andnumerical simulations [19,20]): ‘‘to determine the behavior of

Page 2: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

152 D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

non-Newtonian liquids in a number of simple flow situation usingsuitable defined material functions’’ [18].

To fulfill this objective it is necessary not only to find general/universal solutions for the equation of motion, but also to imposeproper boundary conditions for the working domain. However, inrheometry (especially for complex fluids) it is difficult to alwaysput well posed boundary conditions, since the viscoelastic solu-tions might need ‘‘more boundary conditions than are suppliedby the no-slip condition’’ [18].

In rotational rheometry (in particular plate–plate configuration)the boundary conditions are directly related to the fluid behaviorat the walls (adherence or slip) and to the shape of the free surfaceat the edge of the plates (which is aprioric unknown). If the no-slipcondition is normally accepted at the plates (considered to be per-fectly smooth surfaces), it is almost impossible to control theboundary condition at the edge of the geometry.

Ken Walters inferred the importance of boundary conditions inrheometry and their direct connection to the quantification andinterpretation of the experimental errors; in consequence, eachof the chapters from Rheometry includes a paragraph dedicatedto the analysis and discussion of possible sources of errors.

In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge andend effects, respectively (ii) instrument imperfections. Both arerelated with the fact that boundary of the flow domain is not pre-cisely defined, and this is not referring only to the unknown freesurface at the edge of the plates, but also with the deviation ofthe real geometry of the plates from the calculus geometry (dueto misalignment, lack of parallelism, tilted axes).

We shall refer to these errors as end/edge effects and gap errors;we suppose that no-slip condition to solid walls holds and the freesurface between the plates is cylindrical and in contact with theatmospheric pressure. End effects include inertia influence in afinite geometry (onset of secondary flows) and the error inducedby the approximation of the real free surface of the sample at theedge with a cylindrical surface. Gap error is considered to be gen-erated mainly by the non-parallelism of the plates, so the real gapis not constant along the surfaces. Since the goal of the work is toinvestigate the influence of pattern plates on the measurements inplate–plate geometry, it is compulsory to evaluate first the contri-bution of the end effects and gap error on the experiments per-formed for the smooth (commercial) plates, see Refs. [21–23].

The structure of the paper is the following: Section 2 is dedi-cated to the characterization of fluid samples and the measure-ments of the torques in simple shear for the smooth andpatterned surfaces. A presentation of the surfaces microgeometryis also made. In Section 3 the gap error is analyzed in relation toa thin film (lubrication) analytical Stokes solution for rotationalnon-parallel surfaces. The results from numerical simulations ofNewtonian and generalized Newtonian fluids (Carreau model) inplate–plate smooth and patterned configurations are shown in Sec-tion 4. Finally, in Section 5 the experimental and numerical resultsare analyzed and the conclusions of the work are presented.

2. Experimental

The investigations are performed with the Anton Paar PhysicaMC301 rheometer in controlled strain mode using the parallelplates configuration with diameters of 25 mm, respectively50 mm, at constant temperature within a range from 10 �C to25 �C. The reference values for the samples shear rheology areobtained using the cone and plate configuration (cone diameterof 50 mm and 1� angle). In experiments the upper plate was alwaysthe regular commercial stainless steel plate. Several lower plates ofdifferent materials and patterns have been tested: (i) current lower

plate of the rheometer (PN-plate), (ii) perfectly smooth siliconplate (Si-plate), (iii) silicon plate with pillars pattern (Si-pillars),(iii) copper alloy plate with microchannels pattern (channelsplate).

Silicon wafers of 76 mm in diameter were processed at the IMTBucharest (National Institute for Research and Development inMicrotechnologies) using photolithography and DRIE (Deep Reac-tive Ion Etching) techniques to produce surfaces with pillars pat-tern, see also Ref. [24]. The plates patterned with parallelmicrochannels were obtained by classical mechanical procedure,see Fig. 1.

The most difficult part of the experimental protocol was toobtain a working set-up with minimum alignment and parallelismerrors of the lower plates, relative to the rotational upper geometryplate. First, the horizontal position of the original lower plate wasadjusted and the calibration to some prescribed gaps between theplates was performed using the measurements in three points ofnormal force between the upper plates and the feeler gauges of50 lm and 100 lm (nominal height) mounted on the lower plate.The zero gap corresponds to the position where the measured nor-mal (axial) force against the upper plate is zero. Therefore, in ourcase the prescribed nominal gap of the rheometer indicates actu-ally the lowest gap height between the plates.

Each manufactured lower plate is fixed on the rheometerlower plate and the described calibration was performed beforestarting a new measurement. However, this procedure did noteliminate the lack of parallelism between the tools and the mis-aligned plate–plate problem generated by the gap variation alongthe contact surfaces. Using a set of feeler gauges of 40 lm,50 lm, 60 lm, 70 lm, respectively 150 lm and 160 lm, we mea-sured a gap difference between opposite edges of the 25 mmplate diameter of approximate 10 lm (for all nominal gapsmagnitude).

The samples used in experiments are two Newtonian liquids:Si-oil (silicone oil, with nominal viscosity of 0.4 Pa s at 20 �C) andEn-oil (10W50 engine oil, with nominal viscosity of 0.4 Pa s at10 �C and 0.275 Pa s at 20 �C), and a PIB-solution, a weakly elasticpolymer solution of polyisobutylene with Mw = 0.5 mil. (fromSigma Aldrich) in En-oil, with zero shear viscosity of 1.55 Pa s at10 �C and 0.9 Pa s at 20 �C.

The reference temperature for each test and fluid was fixed asfunction of the measured temperature of the patterned lower plate.This temperature cannot be strictly controlled by the Peltier sys-tem of the rheometer, in consequence the shown data are obtainedat different temperatures.

The oscillatory shear test of PIB-solution is presented in Fig. 2.The measured data disclose two phenomena: (1) the decreasingof the measured viscosity with reducing the gap in plate–plate con-figuration (so called the gap error effect), and (2) the influence ofthe lower plate quality on the measured torque. The first phenom-ena is well known in simple shear rheology and was recently inves-tigated and analyzed in relation to the shear rheometry at highrates [25–28], see also Refs. [21–23]. The influence of the platequality (assuming the plate is smooth) is determined by the adh-erece of the fluid at the material surface and possible slip occu-rence. In Fig. 2 the measured differences in complex viscositybetween the PN-plate and Si-plate are up to 20%, for the samevalue of the gap.

The samples (Newtonian liquids and PIB-solution) are notexpected to exhibit significant slip at the walls of commercialplates. However, the measurements performed with perfectsmooth Si-plates indicate possible presence of slip for PIB-solutionin plate–plate geometry, the phenomenon which seems to beabsent in the cone–plate configuration (where the recorded dataare in the range of the experimental errors for the two testedsmooth lower plates).

Page 3: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

(a)

(b)

(c) Fixed Si-pillars lower plate Fixed channels lower plate

Smooth upper moving plate

240 µm

300 µm300 µm

11.25 µm 8.6 µm

9 µm

φ

Fig. 1. The patterned plates used in experiments: (a) silicon pillars plate with uniform distribution of cylindrical pillars on the surface (by courtesy of IMT Bucharest), (b)channels plate (by courtesy of TU Darmstadt). The average dimensions of micro-geometries and the working setups are also shown (c).

10-1 100 101 1020,60

0,65

0,70

0,75

0,80

0,85

0,90

0,95

Com

plex

vis

cosi

ty [P

as]

CP Si-plate , CP PN gap [µm] 140 180Si-plate , PN ,

(a) (b)

10-1 100 101 102

10-4

10-3

10-2

10-1

100

101

102

G' G'' PN Si-plate PN Si-plate CP , , , gap [µm]140 , , , 180 , , ,

1:1

Dyn

amic

s m

odul

i [Pa

]

PIB-solution

1.6:1

ω [rad/s] ω [rad/s]

Fig. 2. The dependence of the complex viscosity (a) and the dynamics moduli of the PIB-solution (b) on the gap height and the quality of the smooth lower plate. Testesperformed in oscillatory controlled strain frequency sweep at 0.1 [�] strain amplitude at 20 �C, with upper plate diameter of 50 mm. The reference measurements areobtained with cone and plate geometry (CP); the slopes of increasing G0 and G00 with frequency x are shown.

D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162 153

The gap error effect in torsional (plate–plate) flow is associatedto the decreasing of the measured viscosity as the gap height h isdecreasing. This phenomenon is generated by the geometrical non-conformity of the working configuration in comparison to the idealgeometry used in computations. In particular, the error is mainlyinduced by the non-parallelism of the plates. In this case the gapheight is not homogeneous between the plates (as we have alreadyreported). The zero gap procedure indicates the lowest value of thegap height, same value being used to command the rim shear rate.In consequence, the measured torque Tm will be less than the the-oretical torque Tt because the real gap is actually larger than theimposed one.

The theoretical torque in plate–plate configuration is consid-ered to be identical with the corresponding torque in a steady tor-sional flow of an equivalent Newtonian fluid with viscosity g0,

Tt ¼pxR4

2hg0; ð1Þ

where x is the relative angular speed between the plates, R is theradius of the plates and h is the constant height of the gap.

In experiments, the measured shear stress rR and the com-manded shear rate _cR are calculated with the relations

rR ¼2Tm

pR3 ; _cR ¼xRh; ð2Þ

the computed viscosity function being defined as:

g :¼ rR

_cR: ð3Þ

The formula of the shear rate in Eq. (2)2 is valid only for per-fectly smooth parallel plates. For non-parallel discs or patternedplates the value of _cR represents an apparent shear rate becausethe local flow is no more a homogeneous pure shear in the gap(especially for lower patterned surfaces).

Relations (1) and (2) are valid for no-slip boundary conditionsand do not take into consideration neither the end-effects (the

Page 4: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

154 D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

plates are considered infinite), the inertia influence (Reynoldsnumber tends to zero) nor the Weissenberg correction [13,18].

We have to mention that Weissenberg correction (representedby the term @ ln rR=@ ln _cR applied to (2)1 is not related to the gapeffect, but with the non-linearity of the viscosity function (if@ ln rR=@ ln _cR ¼ 1 the formula (2)1 is correct), see Fig. 3.

The gap error effect, induced by the lack of parallelism betweenthe plates, on the computed viscosity (3) becomes more relevant asthe gap is smaller and the fluid viscosity is lower. Some other phe-nomena which may influence the calculation of viscosity (the end-effects, inertia, Weissenberg correction for the shear stress at therim in the non-linear viscoelastic regime) cannot explain therecorded relatively large differences between the measured viscos-ity at different gaps in the plate–plate configuration, especially ath < 0.2 mm, see for details Refs. [21,23].

The influences of gap and patterned surface in simple shear ofPIB-solution and En-oil for a Si-pillars and channel plate patternedsurfaces are presented in Figs. 4 and 5 (the effect of microgeometryon the measurements in plate–plate rheometry was previouslyreported by the authors in Ref. [29]).

All performed experiments in shear (different samples, temper-atures and shear rates) confirm that, at the same apparent shearrate, the measured viscosity is decreasing by decreasing the gapheight. We also recorded lower values of the measured viscosityfor patterned plate as compared to the values measured viscosityfor smooth (normal) plate, at the same gap and same apparentshear rate.

The difference between smooth and patterned surfaces are alsonoticed during stress relaxation of PIB-solution. For smooth platesstress relaxes almost immediately to zero at any gap; for patternedplates the relaxed shear stress reaches fast a small but constantresidual stress, its value being increased by decreasing the gap,see Fig. 6.

The experimental investigations in shear are summarized inFig. 7, where the measured viscosity as function of gap is repre-sented for different lower plate quality and microgeometries atthe apparent shear rate of 50 s�1.

3. Gap influence

The gap error in plate–plate geometry is remarkable for nomi-nal gaps below about 200 lm. The phenomenon was probablythe first time investigated in Ref. [21] and in the last decade severalstudies on the gap error were published by Stokes et al., e.g. Refs.[22,23,25], in the most recent one being proposed a GLM

101 1020,25

0,30

0,35

0,40

0,45

Mea

sure

d vi

scos

ity [P

as]

Shear rate [1/s]

50µm , 100µm , 200µm 300µm , 400µm , 500µmCP ; 0= 0.4 Pas

(a)η

Fig. 3. The gap error effect in shear: (a) measured viscosity as function of the gap magnituplate diameter of 50 mm, 20 �C).

(generalized linear model) methodology to analyze the experimen-tal data and estimate the gap error [23]. The subject has become ofinterest especially in relation with the novel procedures to mea-sure rheological properties at high and very high shear rates basedon the microgap rheometry, see Refs. [25–28].

In classical shear rheometry is generally accepted to representthe gap error by dimension he defined as:

he ¼ hg0

gmðhÞ� 1

� �; ð4Þ

where h is the prescribed gap (the input value for the rheometer,also the value used to compute the apparent shear rate _cR) andgm(h) is the measured viscosity (3) for the imposed shear stress(2)1 [21,23]. The gap errors (4) are calculated for our data and thevalues he(h) are shown in Fig. 8.

In Fig. 8 the reference value for g0 is given by the measurementin cone-plate configuration. In all cases, within an experimentalerror of ±2%, g0 almost coincides with the vicosity measured inplate–plate geometry for 300 lm gap.

The calculated dimension he in (4) does not really represent thetrue error in the measurement of the gap; it is just a ‘‘virtualdimension’’ added to h in order to calculate accordingly to (3)the correct viscosity (at given velocity of the upper plate and themeasured rR).

In the case of a fluid with constant viscosity, no-slip conditionsand smooth plates surfaces, in the value of he include 3 types oferrors: (i) gap error (lack of plates parallelism/misalignment,which is considered to be dominant), (ii) end/edge effect (the realedge surface of the sample is not cylindrical), (iii) inertia/presenceof secondary flows (Re number influence).

The effect of patterned surfaces observed in our tests (see Figs. 7and 8, for example) is not generated by the one of these phenom-ena. The influence of surface quality on the measurements is inves-tigated and analyzed in the next paragraph using the results ofnumerical simulations of the flow field in the gap.

In the absence of the complete 3D topography of the plates andaccurate 3D measurements of plates misalignments, the gap errorcan be explained, and to some extend even computed, using thethin film theory [30,31].

The velocity distribution in a thin film of a Newtonian fluidbetween two non-parallel discs of radius R and gap h� R, seeFig. 10, is approximated by the relation,

v ¼ 12g0r

zðz� hÞ @p@uþ rx

h� zh

; ð5Þ

-1 0 1 2 3 4 5 6 7

-6

-5

-4

-3

-2

-1

0

1

2

Log

[Tor

que]

Log [Shear rate]

decreasing the gap

1:1

(b)

de; (b) the Weissenberg correction coefficient is equal to one (Si-oil, PN-plate, upper

Page 5: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

30 40 50 60 70 80 90 100 110 1200,6

0,7

0,8

0,9

1,0

1,1

1,2

1,3

500 [1/s]

50 [1/s]

Si-pillars plate

Visc

osity

[Pas

]

Time [s]

25 µm 50 µm 75 µm 100 µm

5 [1/s]

(a)

30 40 50 60 70 80 90 100 110 120

0,15

0,20

0,25

0,30

0,35

500 [1/s]50 [1/s]

25 µm 50 µm 75 µm 100 µm

Si-pillars plate

Visc

osity

[Pas

]

Time [s]

5 [1/s](b)

Fig. 4. The gap influence; measured viscosity at constant apparent shear rate (5 s�1, 50 s�1, 500 s�1, respectively) for PIB-solution (a) and En-oil (b) at 10 �C (Si-pillars patternlower plate; upper plate diameter of 25 mm).

30 40 50 60 70 80 90 100 110 1200,3

0,4

0,5

0,6

0,7

0,8

0,9

1,0

1,1

1,2

1,3

500 [1/s]

50 [1/s]

Visc

osity

[Pas

]

Time [s]

gap [μm] PN Channels 25 50 75 100

5 [1/s](a)

30 40 50 60 70 80 90 100 110 1200,1

0,2

0,3

0,4

500 [1/s]

50 [1/s]

Visc

osity

[Pas

]

Time [s]

gap [μm] PN Channels 25 50 75 100 5 [1/s]

(b)

Fig. 5. Patterned surface and gap influences; measured viscosity at constant apparent shear rate (5 s�1, 50 s�1, 500 s�1, respectively) for PIB-solution (a) and En-oil (b) at15 �C; comparison between the normal lower plate and the channels pattern lower plate (upper plate diameter of 25 mm).

0 100 20020

2530

3540

4550

55

6065

7075

80

100 105 110 115 120 125 130 135 140 145 15010-5

10-4

10-3

10-2

10-1

100

Shea

r stre

ss [P

a]

Time [s]

200 [µm ]: , 100 [µm ]: 50 [µm ]: , 25 [µm ]:

PIB - RELAXATIONShea

r stre

ss [P

a]

Time [s]

h [µm ]: 300 200 100 50 25 Channels Plate: , , , , PN - Normal Plate: , , , ,

Fig. 6. The patterned surface and gap influences; shear stress vs. time followed byrelaxation for PIB-solution at _c ¼ 50 s�1 (temperature 10 �C, upper plate of 25 mm).The lower limit of the torque transducer corresponds to the shear stress value of5 mPa, below this value some oscillations of the shear stress are observed.

0 25 50 75 100 125 150 175 200 225 250 275 300

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

oil: PN- normal plate oil: channels plate PIB-solution: Si-plate PIB-solution: Si-pillars PIB-solution: PN- normal plate PIB-solution: channels plate

Mea

sure

d vi

scos

ity [P

as]

Gap h [µm]

Plate diameter 25 mmTemperature 10 oC

Fig. 7. The gap and patterned surface influences on the measurement of shearviscosity in plate and plate geometry at constant shear rate _c ¼ 50 s�1 for En-oil andPIB-solution.

D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162 155

Page 6: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

Fig. 8. The gap error at constant shear rate for En-oil and PIB-solution (PN-plate andchannel plate), see Fig. 7. The error he is much larger for the patterned channel platein comparison with the smooth normal plate. Fig. 10. The correction function j(x) and its variation with the gap for different

values of e [lm].

156 D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

where x is the constant relative angular velocity between theplates, v = vu(r, z) is the rotational velocity and g0 is the fluidviscosity.

Relation (5) is directly obtained from the integration of theNavier–Stokes equation in the limit of zero Reynolds number, withvr, vz� vu.

The pressure distribution in the gap is obtained from the inte-gration of the continuity equation, i.e.

R h0 @v=@udz ¼ 0, respectively

as solution of Ricatti equation

p00 þ 3h

p0 þ k

h3 ¼ 0; ð6Þ

where p = p(h), for h 2 [h1, h2] with h ¼ h1 � eu0

u, for u 2 [0, u0],e = h1 � h2 and k ¼ 6g0r2xu0

e , see Fig. 9.Relation (6) is reduced to the equation p0 ¼ C

h3 � kh2, with

C ¼ 2k h1h2h1þh2

, and the relative pressure distribution within the gapis:

p ¼ 6g0xr2u0

e1h� 1

h1 þ h2� 1

h2

h1h2

h1 þ h2

� �; ð7Þ

where p = 0 for h = h1 and h = h2.The shear stress acting on the upper plate is calculated with the

formula,

r ¼ g0@v@z z¼h

¼ g0rxhþ eh

2rh0p0

¼ g0rxh

1þ 32h1h2

hðh1 þ h2Þ� 1

� �� �; ð8Þ

Fig. 9. Non-parallel plate–plate configuration with variable gap h.

and the torque (2)1 is computed by the relation,

T� ¼ u0

e

Z R

0

Z h2

h1

rr2drdh; ð9Þ

respectively, for u0 = 2p and x = e/h2, the shear stress at the rim ofthe plate becomes:

r�R ¼2T�

pR3 ¼ jðxÞ � g0 _cR; ð10Þ

where

jðxÞ ¼ 2 �1x

lnð1þ xÞ þ 32þ x

� �: ð11Þ

The expression (11) is the correction function for the computed/measured viscosity (3), i.e. j(x) = gm/g0, if the two plates are notparallel (j(x) = 1 for x = 0, respectively for h1 = h2), see Fig. 10.We have to remark that in experiments the input nominal gap his given by the values of h2.

Relations (7)–(11) are found in classical books of fluid mechan-ics as particular solution of the Reynolds lubrication equation forslider bearings [32,33], but here the solution is obtained by directintegration of the Newtonian thin film flow in cylindricalcoordinates.

This calculus is qualitatively relevant for our study because itproves that non-parallelism of the plates in rotational rheometer,represented by the value of e, generates a lower measured viscositythan expected, as the nominal gap h = h2 is decreasing. Relation(11) and the diagrams from Fig. 10 might also offer a quantitativecorrection of the measured viscosity gm. This result is explored inthe next sections of the paper, considering the measured valuee = 10 lm for the gap difference (value reported in §2).

4. Numerical simulations

Numerical solutions for the plate–plate geometry, respectivelyplate–channels geometry, were obtained using the commercialFluent code for the Newtonian fluid and Carreau model,

gð _cÞ � g1g0 � g1

¼ 1þ ðk _cÞ2h in�1

2; ð12Þ

where _c is the local shear rate, g0 is zero shear viscosity, g1 is theinfinite viscosity, n is the shear thinning exponent and k is the time

Page 7: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162 157

constant. The Carreau model is one of the most versatile and com-mon relation to represent the shear thinning behavior with a con-tinuous and derivable viscosity function between finite limits (g0,respectively g1).

Fluent is a robust code which solves the Navier–Stokes flowswith constant or variable viscosity using the finite volume tech-nique. The simulations were performed using the steady viscouslaminar solvent with SIMPLE (semi-implicit method) algorithmfor pressure–velocity coupling and second order discretization.The relative convergence error is 10�8 for the equation of continu-ity and velocity components.

The input viscosities for the Newtonian flows wereg0 = 1.06 Pa s and g0 = 0.1 Pa s and the Carreau model was definedby the following constants: g0 = 1.06 Pa s, g1 = 1.0 mPa s, k ¼ 10 sand n = 0.7, 0.2 and �1, respectively.

Solutions are obtained for the plate diameter d = 25 mm at dif-ferent gaps in the range 25 lm 6 h 6 300 lm. The upper plate isrotated with constant speed, corresponding to the shear rate (2)2

in the interval _cR 2 ½10�2 � 103� s�1 and the lower plate is at rest.The rotational plate has always been considered smooth and thelower plate has been taken to be smooth or patterned with chan-nels geometry (see Fig. 1(b and c)). No-slip boundary condition isimposed at the plates surface and the edge of the geometry is con-sidered a cylindrical surface at constant pressure.

The grid was built based on uniform distributed hexahedralsimilar mesh cells for all configurations, so the number of nodesvaries with the gap and type of geometry (normal or channel plate)between 2 and 3 million, with the remark that density of the nodes(respectively cells) was increasing at the edge of the plates. Thechannel geometry was discretized with quad-elements (typemap), respectively tri-elements (pave type) were used to meshthe upper moving plate.

One of the aim of the simulations was to compute the torqueacting on the upper plate of the PN geometry (Tm = Tc) and to com-pare the input viscosity, g0 or viscosity function (12), to the

Table 1Theoretical and computed torques at h = 100 lm and _c ¼ 1s�1.

Theoretical torque Tt (lN m) 2D axial-symmetric (finite)

Tc Tc/Tt

3.25204 3.49107 1.07

Table 2End-effect and gap influences: 2D axial-symmetric and 3D simulations at _c ¼ 1 s�1. The re

h (lm) 200 100 75

2D 3D 2D 3D 2D

Tc/Tt 1.176 1.07 1.07 1.03 1.05Tc/T200 1 1 0.913 0.96 0.893

12.38 mm 12.32

V = 0

(a)

Fig. 11. The end-effect; the spectrum of the computed radial velocity at the edge of thR = 12.5 mm, h = 100 lm and _c ¼ 1 s�1).

calculated viscosity from (3). The results for the Newtonian fluidfrom Tables 1 and 2 correspond to _c ¼ 1 s�1 (the ratio Tc/Tt beingalmost identical for the whole range of simulations).

The results from Table 1 show that end-effect is less importantin 3D than in 2D computations, however the error represented bythe ratio Tc/Tt is decreasing by decreasing the gap, see Table 2. Theend-effect is determined by the imposed boundary condition forpressure, which induces a secondary flow in the vicinity of theplate edge and an increase in the local wall shear stress, seeFig. 11. Hence, the calculated viscosity from numerical simulationsis always higher than the input viscosity.

If the reference value for torque is considered at a precise valueof the gap (e.g. h = 200 lm), then the relative measured torque Tc/T200 (and in consequence, the measured viscosity) is monotonicallydecreasing by the gap magnitude, see Table 2. This phenomenon isnot present in the ‘‘infinite geometries’’ and is a measure of theinfluence of the end/edge effect with reducing the gap. However,this effect is always observed in experiments where the end-effectcannot be avoided.

In reality, the decrease of the measured viscosity by thedecreasing of the gap is more significant than the numerical resultsfrom Table 2 show, due to the gap error effect.

Indeed, the simulations of the rotational flows between non-parallel (oblique) plates for e = 10 lm and x = 0.1 (h = 100 lm),respectively x = 0.2 (h = 50 lm), confirm the analytical results fromrelation (11), see Table 3. The simulations at _c ¼ 10s�1 generatevery similar values for the correction factor, so we are confidentto consider that j is independent on the apparent shear rate.

The wall shear stress (WSS) distributions on the diameters ofthe upper and lower plates are represented in Fig. 12 for perfectparallel plates and tilted (oblique) configuration. In the last casethe non-symmetric stress distribution justifies the lower measuredtorque than in the case of perfect alignment plates.

The analysis of the two effects: (i) end effect (due to secondaryflows), and (ii) gap error (due to the gap non-homogeneity) in rela-

2D axial-symmetric (infinite) 3D geometry

Tc Tc/Tt Tc Tc/Tt

3.25204 1.0 3.35398 1.03

ference value for the torque, T200, is taken a gap of 200 lm.

50 25

3D 2D 3D 2D 3D

1.018 1.034 1.009 1.013 1.0030.945 0.88 0.934 0.86 0.93

5 mm

V = 0R

(b)

e plates: (a) 2D simulation, (b) 3D simulation (Newtonian fluid, smooth geometry,

Page 8: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

Table 3Computed torques for the normal parallel plates TcPN and for the oblique plates Tco; jis the correction factor from (11) and jn is the value calculated from numeric withe = 10 lm (d = 25 mm, apparent shear rate _c ¼ 1 s�1, g0 = 1.06 Pa s).

h (lm) 100 50

TcPN (lN m) 3.354 3.283Tco (lN m) 3.188 2.983j 0.95 0.908jn 0.95 0.9

Fig. 12. Wall shear stress distribution on the diameters of the plates for h1 = 60 mmand h2 = 50 mm, see Fig. 9 and Table 3.

158 D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

tion to the experimental data for smooth and pattern surfaces ispresented in the last section of the paper.

Simulations for the 3D configuration were the lower plate ispatterned with channels, see Fig. 1(b), are performed only for par-allel plates. The results disclose the oscillations of the computedwall shear stress in the gap (due to the presence of the channelson the lower plate) and lower values of WSS than the correspond-

upper plate

median plane

(b)

(a)

Fig. 13. Computed shear stress distribution for the Newtonian fluid (the values vary frommedian plane with a detail (h = 100 lm, _c ¼ 1 s�1, g0 = 1.06 Pa s). (For interpretation of ththis article.)

ing normal smooth configurations for the Newtonian fluid and Car-reau models with positive n – exponent, see Figs. 13–15.

The numerical simulations offer the possibility to get importantdetails not only on the kinematics of secondary flows in the micro-channels, see Figs. 16 and 17, but also on the viscosity distributionin the gap for the shear thinning fluids, Fig. 18. In Figs. 16 and 17some of the trajectories exit from the gap. This phenomenon isan effect of numerical errors due to the imposed boundary condi-tions at the edge of the plates corroborated with the mesh quality(the mesh density being imposed by the available computationcapacity). However, the computed flow rate exiting from the gapis of order of 10�10 lg/s and the transported flow rate has the orderof 1 lg/s, so the influence of this error is very limited and affectsthe kinematics only in the very vicinity of the gap rim.

The quantitative comparison between the computation forsmooth and patterned plates (with no-slip boundary conditions)sustains the concept of the apparent slip at microgrooved walls,which has to be distinguished from the effective slip observed atmicrostructured surfaces [34–36], see also Ref. [37]. The numericalresults and the experimental data from §2 are analyzed in the nextparagraph.

5. Analysis and conclusions

The first step in modeling the torsional flow between patternedsurfaces has to be the correct representation of the flow dynamicsin the corresponding smooth geometry. The problem needs a care-ful analysis when the results of the modeling (analytic or numeric)are compared to the experiments quantitatively. The real configu-ration of the tested geometry might be locally quite different fromthe assumed calculus geometry, and the discrepancy between thetwo configurations is most probably increasing by decreasing thegap. Such discrepancies may induce errors in the calculus of globalparameters, as torque friction, and generate false conclusions onthe fluid behavior at the boundaries of the plates.

Many applications of simple and complex flows in the presenceof patterned surfaces are related to microfluidics and tribology. Themajor goal of the studies is to control the boundary conditions, togenerate slip at the walls [38–40] or to create hydrophobic surfacesfor reducing the drag [41,42]. Several microgeometries (pillars,

blue color to red color, see Fig. 12 for the magnitude of WSS): (a) upper plate, (b)e references to color in this figure legend, the reader is referred to the web version of

Page 9: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

n n

(a)

(b)

Fig. 14. Computed shear stress distribution for the Carreau models (n = 0.2 and n = �1) on the median plane: (a) _c ¼ 1 s�1, (b), _c ¼ 10 s�1 (h = 100 lm).

0,000 0,002 0,004 0,006 0,008 0,010 0,01210-2

10-1

100

101

102

WS

S [P

a]

Radius [m]

Rate [1/s] normal channel 1: , 10: , 100: ,

100

10

1

(a)

0,000 0,002 0,004 0,006 0,008 0,010

10-2

10-1

100

n = –1

n = 0.2W

SS

[Pa]

Radius [m]

n: 1 0.2 –1 normal: , , channel: , , n = 1

Rate = 1 [1/s]

(b)

Fig. 15. Wall shear stress distribution on the diameter of the upper (moving) plate for h = 100 lm: (a) Newtonian fluid (g0 = 1.06 Pa s, apparent shear rate as parameter), (b)Carreau model (n – exponent as parameter).

(a) (b)

(c)

1 s-1 10 s-1 100 s-1

Fig. 16. Flow trajectories in the gap for the channels plate geometry (colored by the velocity magnitude): (a) Newtonian fluid g0 = 1.06 Pa s, (b) Carreau model with n = 0.7(h = 100 lm, _c ¼ 10 s�1); (c) details with the flow trajectories for the Carreau model with n = 0.2 at _c ¼ 1 s�1, _c ¼ 10 s�1 and _c ¼ 100 s�1. (For interpretation of the referencesto color in this figure legend, the reader is referred to the web version of this article.)

D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162 159

Page 10: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

Newtonian Carreau (n = 0.2)

Carreau (n = 0.7) Carreau (n = - 1)

Fig. 17. Details of the secondary flow patterns in the channels of the lower plate for the Newtonian and the Carreau models ( _c ¼ 1 s�1).

Fig. 18. Viscosity distribution in a channel for the Carreau model (n – exponent as parameter). Lower viscosity corresponds to the surface of the upper (moving) plate; shearbanding phenomena is observed for negative n-exponents (h = 100 lm, _c ¼ 1 s�1, _c ¼ 10 s�1).

160 D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

microgrooves, channels, strips) have been tested for those pur-poses. Since it is not easy to obtain a complete description of theflow kinematics in the vicinity of such surfaces, the target of thestudies is to evaluate the hydrophobicity of a given surface patternas function of the measured slip velocity.

But the observed slip velocity is not always the lack of adher-ence of the fluid to the solid wall. In many applications the lackof adherence is only apparent, the phenomenon being generatedby wall depletion, shear banding or the presence of multiphaseflows (i.e. the existence of the air in microgrooves).

The main aim of our work was to evaluate the contribution ofthe patterned surface on the measurements in plate–plate geome-try and to give a quantitative representation of the experimentalresults. Firstly we have analyzed the contribution of the end effectsand the gap error for smooth plates under no-slip boundary condi-tions. If the correction factor j (which represent the contribution

of the gap error, see Table 3) is corroborated with the end-edgeeffect contribution from Table 2, one can compute the total relativemeasured torque – (Tc/T200)T for the experiments performed in realconditions, e.g. (Tc/T200)T = 0.90�0.934 ffi 0.84 for nominal gap ofh = 50 lm and e = 10 lm. Consequently the measured viscosity ath = 50 lm will be only 84% of the viscosity recorded at the gap of200 lm. We notice that this result is not determined by any slipor lack of adherence at the plates surfaces.

Real or apparent slip might generate at macroscopic scale sim-ilar consequences with the non-parallelism of the plates (gaperror) or with the presence of the micro-structures at the wall.The decreasing of torque in plate–plate rheometry for patternedsurfaces in comparison with the smooth surface is such conse-quence, see Fig. 7. In Table 4 computed torques are presented atthe upper plate (based on numerical simulations with no-slipboundary condition) for normal and channel lower plates (TcPN,

Page 11: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162 161

respectively TcCh) and also the relative torque aT = TcCh/TcPN. There-fore, at the gap of 50 lm, and the same apparent rate, we expect adecreasing in friction on the upper plate by 21% for patternedlower plate, phenomenon which is determined by the change inthe flow kinematics within the gap, see Figs. 16 and 17.

A sketch of the simple shear flow in the vicinity of a smooth andpatterned surface is represented in Fig. 19. Here b is the real sliplength, with b⁄ = b + H, where H and d are the characteristic dimen-sion of the microgeometry. The gap h and the applied velocity V�0on the upper plate is the same in both cases. We assume that atsmooth surfaces the no-slip boundary condition is applied.

In the case of patterned plate (Fig. 19.b) we might have real(effective) slip at the wall (slip velocity Vsw), slip at y = H (slip veloc-ity Vs) or apparent slip (velocity Vas at y = d, d > H, with zero velocityat y = H), which is associated to a high velocity gradient in thevicinity of the wall. In all cases the velocity distribution disclosesthe same value at y = d and identical patterns for y > d, so the realslip cannot be easily differentiated from the apparent slip.

One procedure to analyze the existence of slip at the wall isbased on the definition of the structural parameter b⁄,

b� ¼ r�0r�s� 1

� �h; ð13Þ

where r�0 is the measured shear stress at the upper plate of asmooth geometry and r�s is the measured shear stress at the upperplate of a geometry with patterned lower wall [29], see also Refs.[34,35].

For a given microgeometry and r�0 > r�s , the computation of b⁄

might characterize the flow regime on patterned surface relative tothe flow at the smooth surface. A positive dimension br: = b⁄ � Hdefines a flow regime with effective induced slip at the patternedsurface; if br < 0 the flow is associated to the apparent slip inducedby the plate’s microgeometry.

In the case of numerical simulations performed for the pat-terned channels pattern with H = 240 lm, the value of br is nega-

Table 4Computed torques for normal plate and channel plate (d = 25 mm, apparent shearrate _c ¼ 1 s�1, g0 = 1.06 Pa s).

h (lm) 100 75 50 25

TcPN (lN m) 3.354 3.31 3.283 3.261TcCh (lN m) 2.928 2.782 2.595 2.288aT 0.87 0.84 0.79 0.7b⁄ 14.94 14.28 13.29 10.71

Fig. 19. Simple shear flow in vicinity of smooth and pattern surfaces.

tive at all gaps, b⁄� H, see Table 4. Therefore the computed flowregime is characterized by apparent slip on the patterned surface.

Fig. 20 shows the computed structural parameter b⁄ for theexperimental data from Fig. 7. Within the frame of our interpreta-tion, the effective induced slip is present only for the Si-plate pil-lars pattern (H = 9 lm, see Fig. 1a), and remarkable only for theNewtonian oil with low viscosity.

This result is consistent with the statement that induced slip atthe patterned walls is effective for microgeometries with aspectratio around one (0.5 < H/d < 1.5) for height dimension H in therange of microns or tens of microns (see Figs. 19 and 1).

In the case of channel pattern the aspect ratio is around one, butthe H value is large, i.e. H = 240 lm. In this case the microgeometryinduces an apparent slip and not an effective slip, since b⁄� H forboth samples, see Fig. 20.

This conclusion is confirmed by the results from Fig. 21. Here,the results of computations are superimposed for the smoothand patterned channels geometries (end/edge effects, gap errorcorrection, respectively the patterned surface effect) and comparedwith the measured values (in this representation the assumptionthat the relative computed correction coefficients have the samevalues in the range of tested shear rates and viscosities has beenconsidered). The difference between the measurements and thecomputations are in the range of 10% for the investigated domain.

Fig. 20. Variation of parameter b⁄ with the gap.

Fig. 21. Computed vs. experimental relative viscosity gm/g0 for h 6 100 lm (g0

corresponds at h = 200 lm).

Page 12: Journal of Non-Newtonian Fluid Mechanics JNNFM 2015.pdf · In the present paper we are mainly concerned with two catego-ries of errors mentioned by Ken Walters in Ref. [18]: (i) edge

162 D. Broboana et al. / Journal of Non-Newtonian Fluid Mechanics 222 (2015) 151–162

The present study demonstrates that hydrophobic effects can beinduced to the walls, without violating of the no-slip condition, bychanging the local flow spectrum due to the presence of patternedsurfaces at the solid boundary. The results confirm that computa-tional rheometry is an useful tool not only to interpret the exper-imental data and to calculate the errors of the measurements,but also to explore and model flow phenomenon as apparent slip.

In particular, the work was dedicated to the investigation of theinfluence of patterned surfaces on the measured torque in plate–plate configuration. The correlations of the experimental data withnumerical simulations reveal the contribution of the patterned sur-face on the flow dynamics within the gap and the possibility tomake the distinction between the effective and the apparent slipinduced by the microgeometry. The paper presented only numeri-cal solutions for pure viscous and Carreau models. Of course, thecontinuation of the work has to include the effects of elasticity,but this approach is at the moment beyond the computationalcapabilities available in our group.

This study is considered to be of interest for developing noveltechniques in rheometry (using for tests plates with well definedmicro-patterned surface) and in microfluidics, where the controlof boundary conditions and wall adherence are crucial for manyapplications.

Acknowledgements

Corneliu Balan acknowledges the continuous support in hisactivity offered since 1991 with generosity by Professor Ken Walt-ers FRS.

The authors acknowledge Dr. ing. Catalin Marculescu (IMTBucharest) and Ph.D. student Amon Klausmann (T.U. Darmstadt)for their assistance in fabrication of the patterned surfaces, andalso the financial support from the grants of the Ministry ofNational Education, Romania, CNCS-UEFISCDI: PN-II-ID-PCE-2012-4-0245 and PN-II-PT-PCCA-2011-3. The work of NicoletaOctavia Tanase has been funded by the Sectorial Operational Pro-gramme Human Resources Development 2007–2013 of the Minis-try of European Funds through the Financial Agreement POSDRU/159/1.5/S/132395.

References

[1] X. Ding, P. Li, Sz.-C.S. Lin, Z.S. Stratton, N. Nama, F. Guo, D. Slotcavage, X. Mao, J.Shi, F. Costanzoa, T.J. Huang, Surface acoustic wave microfluidics, Lab Chip 13(2013) 3626–3649.

[2] A.D. van der Meer, V.V. Orlova, P. ten Dijke, A. van den Berga, Ch.L. Mummeryd,Three-dimensional co-cultures of human endothelial cells and embryonic stemcell-derived pericytes inside a microfluidic device, Lab Chip 13 (2013) 3562–3568.

[3] A.F. Chrimes, K. Khoshmanesh, P.R. Stoddart, A. Mitchella, K. Kalantar-zadeha,Microfluidics and Raman microscopy: current applications and futurechallenges, Chem. Soc. Rev. 42 (2013) 5880–5906.

[4] E. Lauga, M.P. Brenner, H.A. Stone, The no-slip boundary condition, in:Handbook of Experimental Fluid Dynamics, Springer, Berlin, 2007, pp. 1219–1240.

[5] T. Sochi, Slip at fluid-solid interface, Polym. Rev. 51 (2011) 309–340.[6] J.P. Rothstein, Slip on superhydrophobic surfaces, Annu. Rev. Fluid Mech. 42

(2010) 89–109.[7] D.C. Tretheway, C.D. Meinhart, Apparent fluid slip at hydrophobic micro-

channel walls, Phys. Fluids 14 (2002) L9–L12.[8] S. Heidenreich, P. Ilg, S. Hess, Boundary conditions for fluids with internal

orientational degrees of freedom: apparent velocity slip associated with themolecular alignment, Phys. Rev. E 75 (2007) 066302.

[9] N.V. Prieyjev, A.A. Darhuber, S.M. Troian, Slip behavior in liquid films onsurfaces of pattered wettability: comparison between continuum andmolecular dynamics simulations, Phys. Rev. E71 (2005) 041608.

[10] N.V. Priezjev, S.M. Troian, Influence of periodic wall roughness on the slipbehaviour at liquid/solid interfaces: molecular-scale simulations versuscontinuum predictions, J. Fluid Mech. 554 (2006) 25–46.

[11] O.I. Vinogradova, Slippage of water over hydrophobic surfaces, Int. J. Miner.Process. 56 (1999) 31–60.

[12] S. Granick, Y. Zhu, H.J. Lee, Slippery questions about complex fluids flowingpast solids, Nat. Mater. 2 (2003) 221–227.

[13] B.D. Coleman, H. Markovitz, W. Noll, Viscometric Flows of Non-NewtonianFluids – Theory and Experiments, Springer, New York, 1966.

[14] Y.M. Joshi, A.K. Lele, R.A. Mashelkar, Slipping fluids: a unified transientnetwork model, J. Non-Newtonian Fluid Mech. 89 (2000) 303–335.

[15] R. Buscall, Letter to the editor: wall slip in dispersion rheometry, J. Rheol. 54(2010) 1177–1183.

[16] A. Yoshimura, R.K. Prudhomme, Wall slip corrections for Couette and paralleldisk viscometers, J. Rheol. 32 (1988) 53–67.

[17] H.A. Barnes, A review of the slip (wall depletion) of polymer solutions,emulsions and particle suspensions in viscometers: its cause, character andcure, J. Non-Newtonian Fluid Mech. 56 (1995) 221–251.

[18] K. Walters, Rheometry, Chapman and Hall, London, 1976.[19] H.A. Barnes, J.F. Hutton, K. Walters, An Introduction to Rheology, Elsevier,

Amsterdam, 1989.[20] M.J. Crochet, A.R. Davies, K. Walters, Numerical Simulation of Non-Newtonian

Flow, Elsevier, Amsterdam, 1984.[21] J. Kramer, J.T. Uhl, R.K. Prudhomme, Measurement of the viscosity of guar gum

solutions to 50,000 s�1 using a parallel plate rheometer, Polym. Eng. Sci. 27(1987) 598–602.

[22] G.A. Davies, J.R. Stokes, On the gap error in parallel plate rheometry that arisesfrom the presence of air when zeroing the gap, J. Rheol. 49 (2005) 919–922.

[23] O. Kravchuk, J.R. Stokes, Review of algorithms for estimating the gap errorcorrection in narrow gap parallel plate rheology, J. Rheol. 57 (2013) 365–375.

[24] X. Li, H. Tian, J. Shao, Y. Ding, H. Liu, Electrically modulated microtransfermolding for fabrication of micropillar arrays with spatially varying heights,Langmuir 29 (2013) 1351–1355.

[25] Ch. Clasen, G.H. McKinley, Gap-dependent microrheometry of complex liquids,J. Non-Newtonian Fluid Mech. 124 (2004) 1–10.

[26] Ch. Clasen, B.P. Gearing, G.H. McKinley, The flexure-based microgap rheometer(FMR), J. Rheol. 50 (2006) 883–905.

[27] G.A. Davies, J.R. Stokes, Thin film and high shear rheology of multiphasecomplex fluids, J. Non-Newtonian Fluid Mech. 148 (2008) 73–87.

[28] S.J. Baik, P. Moldenaers, Ch. Clasen, A sliding plate microgap for thesimultaneous measurement of shear stress and first normal stress difference,Rev. Sci. Instrum. 82 (2011) 035121.

[29] C. Balan, D. Broboana, E. Ionescu, R. Riedel, Influence of hydrophobic surfaceson the measured rheological properties of complex fluids, in: The XVth Int.Congress Rheology, Lisbon, Portugal, August 2012 (pp. 434, book of abstracts).

[30] Ch. Clasen, H.P. Kavehpour, G.H. McKinley, Bridging tribology andmicrorheology of thin films, Appl. Rheol. 20 (2010) 45049.

[31] Ch. Clasen, High shear rheometry using hydrodynamic lubrication flows, J.Rheol. 57 (2013) 197–221.

[32] D.J. Acheson, Elementary Fluid Mechanics, Clarendon Press, Oxford, 1990.[33] F. Durst, Fluid Mechanics: An Introduction to the Theory of Fluid Flows,

Springer, Berlin, 2008.[34] M.Z. Bazant, O.I. Vinogradova, Tensorial hydrodynamic slip, J. Fluid Mech. 613

(2008) 125–134.[35] K. Kamrin, M.Z. Bazant, H.A. Stone, Effective slip boundary conditions for

arbitrary periodic surfaces: the surface mobility tensor, J. Fluid Mech. 658(2010) 409–437.

[36] E.S. Asmolov, J. Zhou, F. Schmid, O.I. Vinogradova, Effective slip-length tensorfor a flow over weakly slipping stripes, Phys. Rev. E 88 (2013) 023004.

[37] C. Schönecker, T. Baier, S. Hardt, Influence of the enclosed fluid on the flowover a microstructured surface in the Cassie state, J. Fluid Mech. 740 (2014)168–195.

[38] M. Tauviqirrahman, R. Ismail, J. Jamari, D.J. Schipper, A study of surfacetexturing and boundary slip on improving the load support of lubricatedparallel sliding contacts, Acta Mech. 224 (2013) 365–381.

[39] G. Daschiel, M. Peric, J. Jovanovic, A. Delgado, The holy grail of microfluidics:sub-laminar drag by layout of periodically embedded microgrooves,Microfluid Nanofluid, http://dx.doi.org/10.1007/s10404-013-1182-0.

[40] D.J. Lee, K.Y. Cho, S. Jang, Y.S. Song, J.R. Youn, Liquid slip on a nanostructuredsurface, Langmuir 28 (2012) 10488–10494.

[41] Y. Xue, S. Chu, P. Lv, H. Duan, Importance of hierarchical structures in wettingstability on submersed superhydrophobic surfaces, Langmuir 28 (2012) 9440–9450.

[42] C. Lee, C-J. Kim, Influence of surface hierarchy of superhydrophobic surfaces onliquid slip, Langmuir 27 (2011) 4243–4248.