Review
A review of the slip (wall depletion) of polymer solutions, emulsions and particle suspensions in viscometers: its cause, character, and cure

https://doi.org/10.1016/0377-0257(94)01282-MGet rights and content

Abstract

Slip occurs in the flow of two-phase systems because of the displacement of the disperse phase away from solid boundaries. This arises from steric, hydrodynamic, viscoelastic and chemical forces and constraints acting on the disperse phase immediately adjacent to the walls. The enrichment of the boundary near the wall with the continuous (and usually low-viscosity) phase means that any flow of the fluid over the boundary is easier because of the lubrication effect. Because this effect is usually confined to a very narrow layer — with typical thickness of 0.1–10 μm—it so resembles the slip of solids over surfaces that it has historically been given the same terminology. The restoring force for all the forces that cause an increase in concentration is usually osmotic, and this will always limit the effective slip. In dilute systems, concentration gradients can be present over relatively large distances out from walls, giving what might be interpreted on an overall basis as a thick solvent-only layer. However, as the concentration of the system increases, the layer gets thinner and thinner because it is more difficult to create with the large reverse osmotic force present. However, the enormous increase in the bulk viscosity with increase in concentration means that although thinner, the layer becomes, paradoxically, even more important.

Slip manifests itself in such a way that viscosity measured in different size geometries gives different answers if calculated the normal way — in particular the apparent viscosity decreases with decrease in geometry size (e.g. tube radius). Also, in single flow curves unexpected lower Newtonian plateaus are sometimes seen, with an apparent yield stress at even lower stresses. Sudden breaks in the flow curve can also be seen.

Large particles as the disperse phase (remember flocs are large particles), with a large dependence of viscosity on the concentration of the dispersed phase are the circumstances which can give slip, especially if coupled with smooth walls and small flow dimensions. The effect is usually greatest at low speeds/flow rates. When the viscometer walls and particles carry like electrostatic charges and the continuous phase is electrically conducted, slip can be assumed.

In many cases we need to characterise the slip effects seen in viscometers because they will also be seen in flow in smooth pipes and condults in manufacturing plants. This is usually done by relating the wall shear stress to a slip velocity using a power-law relationship. When the bulk flow has also been characterized, the flow in real situations can be calculated. To characterise slip, it is necessary to change the size of the geometry, and the results extrapolated to very large size to extract unambigouos bulk-flow and slip data respectively. A number of mathematical manipulations are necessary to retrieve these data.

We can make attempts to eliminate slip by altering the physical or chemical character of the walls. This is usually done physically by roughening or profiling, but in the extreme, a vane can be used. This latter geometry has the advantage of being easy to make and clean. In either case—by extrapolation or elimination—we end up with the bulk flow properties. This is important in situations where we are trying to understand the microstructure/flow interactions.

References (107)

  • OldroydJ.G.
  • MeetenG.H. et al.
  • GarnerF.H. et al.

    Nature (LOndon)

    (1946)
  • AubertJ.M. et al.

    J. Chem. Phys.

    (1980)
  • PetruccioneF. et al.

    J. Rheol.

    (1988)
  • VergnesB. et al.
  • GuidoS. et al.

    Rheol. Acta

    (1994)
  • SorbieK.S.

    SPE

    (1989)
  • ChaffeyC. et al.

    Rheol. Acta

    (1965)
  • SmartJ.R. et al.

    Phys. Fluids

    (1991)
  • A. Hopkins, Personal communication,...
  • WoodcockL.V.

    J. Non-Newtonian Fluid Mech.

    (1986)
  • GoldsmithH.L. et al.

    J. Colloid Interface Sci.

    (1979)
  • LeightonD. et al.

    J. Fluid Mech.

    (1987)
  • Gadala-MariaF. et al.

    J. Rheology

    (1980)
  • HookhamP.
  • UijttewaalS.J. et al.

    Phys. Fluids A

    (1993)
  • ChanP.C.-H. et al.

    J. Fluid Mech.

    (1979)
  • LimF.J. et al.

    J. Rheology

    (1989)
  • DelimeA. et al.

    Rheol. Acta

    (1991)
  • BuscallR. et al.

    J. Rheology

    (1993)
  • ThomasH.W.

    Biorheology

    (1962)
  • MinishitaT. et al.

    Polym. Sci.

    (1969)
  • WhorlowR.

    Rheological Techniques

    (1992)
  • GregoryT. et al.

    Surface Coatings Int (OOCCA)

    (1993)
  • JiZ. et al.

    SPE ANTEC Tech. Papers

    (1990)
  • DzuyN.Q. et al.

    J. Rheology

    (1983)
  • CarnaliJ. et al.

    J. Rheology

    (1990)
  • ReinerM.

    Deformation Strain and Flow

    (1949)
  • CameronJ.J.

    J. Rheology

    (1989)
  • CohenY.
  • YilmazerU. et al.

    Polym. Comp.

    (1991)
  • KalyonD.M. et al.

    J. Rheology

    (1993)
  • BoersmaW.H. et al.

    J. Rheology

    (1991)
  • MooneyM.

    J. Rheology

    (1931)
  • ChengD.C.-H. et al.KiljanskiT.

    Rheol. Acta

    (1989)
  • WeinO. et al.

    J. Rheology

    (1992)
  • YoshimuraA. et al.

    J. Rheology

    (1988)
  • Muller-MohnssenH. et al.

    J. Rheology

    (1990)
  • AgarwalU.S. et al.

    Chem. Eng. Sci.

    (1994)
  • BusseW.F.

    J. Polym. Sci. A-2

    (1967)
  • SchriberH.P. et al.

    J. Polym. Sci.

    (1965)
  • MetznerA.B.

    J. Non-Newtonian Fluid Mech.

    (1979)
  • SekhonG. et al.

    J. Polym. Sci. Polym. Phys. Ed.

    (1982)
  • BrunnP.O.

    Rheol Acta

    (1976)
  • BrunnP.O. et al.

    Rheol Acta

    (1984)
  • BrunnP.O.

    J. Polym. Sci. Polym. Phys. Ed.

    (1985)
  • BrunnP.O.

    J. Rheology

    (1985)
  • BrunnP.O. et al.

    J. Non-Newtonian Fluid Mech.

    (1987)
  • GrisafiS. et al.

    J. Rheol.

    (1989)
  • Cited by (0)

    View full text