Skip to main content
Log in

Rheometry for large-particulated fluids: analysis of the ball measuring system and comparison to debris flow rheometry

  • Original Contribution
  • Published:
Rheologica Acta Aims and scope Submit manuscript

Abstract

For large-particulated fluids encountered in natural debris flow, building materials, and sewage treatment, only a few rheometers exist that allow the determination of yield stress and viscosity. In the present investigation, we focus on the rheometrical analysis of the ball measuring system as a suitable tool to measure the rheology of particulated fluids up to grain sizes of 10 mm. The ball measuring system consists of a sphere that is dragged through a sample volume of approximately 0.5 l. Implemented in a standard rheometer, torques exerted on the sphere and the corresponding rotational speeds are recorded within a wide measuring range. In the second part of this investigation, six rheometric devices to determine flow curve and yield stress of fluids containing large particles with maximum grain sizes of 1 to 25 mm are compared, considering both rheological data and application in practical use. The large-scale rheometer of Coussot and Piau, the building material learning viscometer of Wallevik and Gjorv, and the ball measuring system were used for the flow curve determination and a capillary rheometer, the inclined plane test, and the slump test were used for the yield stress determination. For different coarse and concentrated sediment–water mixtures, the flow curves and the yield stresses agree well, except for the capillary rheometer, which exhibits much larger yield stress values. Differences are also noted in the measuring range of the different devices, as well as for the required sample volume that is crucial for application.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

References

  • Ancey C (2001) Role of lubricated contacts in concentrated polydisperse suspensions. J Rheol 45:1421–1439

    Article  ADS  CAS  Google Scholar 

  • Ancey C, Jorrot H (2001) Yield stress for particle suspension within a clay dispersion. J Rheol 45:297–319

    Article  ADS  CAS  Google Scholar 

  • Ansley RW, Smith TN (1967) Motion of spherical particles in a Bingham plastic. AIChE J 13:1193–1196

    Article  Google Scholar 

  • Atapattu DD, Chhabra RP, Uhlherr PHT (1995) Creeping sphere motion in Herschel–Bulkley fluids. J Non-Newton Fluid Mech 59:245–265

    Article  CAS  Google Scholar 

  • Banfill PFG (1994) Rheological methods for assessing the flow properties of mortar and related materials. Constr Build Mater 8:43–50

    Article  Google Scholar 

  • Banfill PFG, Beaupré D, Chapdelaine F, Larrard F, Domone P, Nachbaur L, Sedran T, Wallevik OH, Wallevik JE (2000) Comparison of concrete rheometers: international tests at LCPC (Nantes, France), NISTIR-Report 6819 of the NIST (National Institute of Standards and Technology)

  • Baudez JC, Chabot F, Coussot P (2002) Rheological interpretation of the slump test. Appl Rheol 12:133–141

    Google Scholar 

  • Beaulne M, Mitsoulis E (1997) Creeping motion of a sphere in tubes filled with Herschel–Bulkley fluids. J Non-Newton Fluid Mech 72:55–71

    Article  CAS  Google Scholar 

  • Beris AN, Tsamopoulos JA, Armstrong RC, Brown RA (1985) Creeping motion of a sphere through a Bingham plastic. J Fluid Mech 158:219–244

    Article  MATH  ADS  CAS  MathSciNet  Google Scholar 

  • Blackery J, Mitsoulis E (1997) Creeping motion of a sphere in tubes filled with a Bingham plastic material. J Non-Newton Fluid Mech 70:59–77

    Article  CAS  Google Scholar 

  • Borgia A, Spera FJ (1990) Error analysis for reducing noisy wide-gap concentric cylinder rheometric data for nonlinear fluids: theory and applications. J Rheol 34:117–136

    Article  MATH  ADS  Google Scholar 

  • Chhabra RP (2007) Bubbles, drop, and particles in non-newtonian fluids. Taylor & Francis, Boca-Raton

    Google Scholar 

  • Chhabra RP, Richardson JF (1999) Non-newtonian flow in the process industries. Butterworth-Heinemann, Oxford

    Google Scholar 

  • Chhabra RP, Uhlherr PHT (1988) Static equilibrium and motion of spheres in viscoplastic liquids. In: Cheremisinoff NP (Ed) Encyclopedia of fluid mechanics, vol 7. Gulf, Houston, pp 611–633

    Google Scholar 

  • Clift R, Grace JR, Weber ME (1978) Bubbles, drops and particles. Academic, New York

    Google Scholar 

  • Coussot P (2007) Plasticity and geophysical flows: a review. J Non-Newton Fluid Mech 142:4–35

    Article  CAS  Google Scholar 

  • Coussot P, Boyer S (1995) Determination of yield stress fluid behaviour from inclined plane test. Rheol Acta 34:534–543

    Article  CAS  Google Scholar 

  • Coussot P, Piau JM (1995) A large-scale field concentric cylinder rheometer for the study of the rheology of natural coarse suspensions. J Rheol 39:105–124

    Article  ADS  CAS  Google Scholar 

  • Coussot P, Proust S, Ancey C (1996) Rheological interpretation of deposits of yield stress fluids. J Non-Newton Fluid Mech 66:55–70

    Article  CAS  Google Scholar 

  • Coussot P, Laigle D, Arattano M, Deganutti A, Marchi L (1998) Direct determination of rheological characteristics of debris flow. J Hydraul Eng (ASCE) 124:865–868

    Article  Google Scholar 

  • Liu KF, Mei CC (1989) Slow spreading of a sheet of Bingham fluid on an inclined plane. J Fluid Mech 207:505–529

    Article  MATH  ADS  CAS  Google Scholar 

  • Major JJ, Pierson T (1992) Debris flow rheology: experimental analysis of fine-grained slurries. Water Resour Res 28:841–857

    Article  ADS  Google Scholar 

  • Metzner AB, Otto RE (1957) Agitation of non-newtonian fluids. AIChE J 3:3–10

    Article  CAS  Google Scholar 

  • Müller M, Tyrach J, Brunn PO (1999) Rheological characterization of machine-applied plasters. ZKG Int 52:252–258

    Google Scholar 

  • Nguyen QD, Boger DV (1987) Characterization of yield stress fluids with concentric cylinder viscometers. Rheol Acta 26:508–515

    Article  Google Scholar 

  • Nguyen QD, Boger DV (1992) Measuring the flow properties of yield stress fluids. Ann Rev Fluid Mech 24:47–88

    Article  ADS  Google Scholar 

  • Parsons JD, Whipple KX, Simoni A (2001) Experimental study of the grain-flow, fluid-mud transition in debris flows. J Geol 109:427–447

    Article  ADS  Google Scholar 

  • Pashias N, Boger DV, Summers J, Glenister DJ (1996) A fifty cent rheometer for yield stress measurement. J Rheol 40:1179–1189

    Article  ADS  CAS  Google Scholar 

  • Phillips CJ, Davies TRH (1991) Determining rheological parameters of debris flow material. Geomorphology 4:101–110

    Article  ADS  Google Scholar 

  • Roussel N, Coussot P (2005) “Fifty-cent rheometer” for yield stress measurements: from slump to spreading flow. J Rheol 49:705–718

    Article  ADS  CAS  Google Scholar 

  • Schatzmann M (2005) Rheometry for large particle fluids and debris flows. Dissertation No 16093, ETH Zurich, Zurich

  • Schatzmann M, Fischer P, Bezzola GR, Minor HE (2003a) The ball measuring system—a new method to determine debris flow rheology? In: Rickenmann D, Chen C (Eds) Proceedings of the 3rd international conference on debris flow hazards mitigation. Millpress, Rotterdam, pp 387–398

  • Schatzmann M, Fischer P, Bezzola GR (2003b) Rheological behaviour of fine and large particle suspensions. J Hydraul Eng (ASCE) 129:796–803

    Article  Google Scholar 

  • Schowalter WR, Christensen G (1998) Toward a rationalization of the slump test for fresh concrete: comparison of calculations and experiments. J Rheol 42:865–870

    Article  ADS  CAS  Google Scholar 

  • Schulze B, Brauns J, Schwalm I (1991) Neuartiges Baustellen-Messgerät zur Bestimmung der Fliessgrenze von Suspensionen. Sonderdruck aus Geotechnik 3/1991. Heidelberger Zement, Berliner Strasse 6, D-6900 Heidelberg

  • Shapley NC, Brown RA, Armstrong RC (2004) Evaluation of particle migration models based on laser Doppler velocimetry measurements in concentrated suspensions. J Rheol 48:255–279

    Article  ADS  CAS  Google Scholar 

  • Tabuteau H, Coussot P, de Bruyn JR (2007) Drag force on a sphere in steady motion through a yield-stress fluid. J Rheol 51:125–137

    Article  ADS  CAS  Google Scholar 

  • Tattersall GH, Banfill PFG (1983) The rheology of fresh concrete. Pitman, London

    Google Scholar 

  • Tattersall GH, Bloomer SJ (1979) Further development of the two-point test for workability and extension of its range. Mag Concr Res 31:202–210

    Article  Google Scholar 

  • Tyrach J (2001) Rheologische charakterisierung von zementären baustoffsystemen. Dissertation, Universität Erlangen-Nürnberg

  • Wallevik OH, Gjorv OE (1990) Development of a coaxial cylinder viscometer for fresh concrete. In: Proceeding of the rilem colloquium. Chapman and Hall, Hanover

    Google Scholar 

  • Whipple KX (1997) Open-Channel flow of Bingham fluids: Applications in debris flow research. J Geol 105:243–262

    Article  ADS  Google Scholar 

  • Wilson SDR, Burgess SL (1998) The steady, spreading flow of a rivulet of mud. J Non-Newton Fluid Mech 79:77–85

    Article  MATH  CAS  Google Scholar 

Download references

Acknowledgement

Specials thanks are given to P. Coussot for his advice regarding the consolidation of the conversion theory in the laminar flow regime with the help of the yield stress criterion.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to P. Fischer.

Appendix: Problems encountered during BMS experiments

Appendix: Problems encountered during BMS experiments

The basic experiment conducted with the BMS consists in measuring the torque T at a given rotational speed Ω while the sphere makes one full rotation. Problems encountered during the measurement of particulated fluids with the BMS are tool acceleration, wake formation and viscous overstream, sphere interaction with suspended particles, data scattering due to grain size and temporary jamming, and the effect of prestructured samples. These problems are typical for many rheometrical measuring techniques and devices, but are considered below.

Tool acceleration

In order to attain a specified rotational speed, the sphere must be accelerated, which requires an additional torque above that required to drag the sphere across a given fluid. Figure 15 shows the start-up flow for samples containing different maximum grain sizes and different sediment concentrations C v. After acceleration, a steady drag flow around the sphere and sphere holder is achieved. In this regime, the measured torque is only dependent on the rheological properties of the fluid. Meaningful rheological data require steady-state flows, which must be established by experiments. As a consequence, only speed and torque data for the steady drag flow regime are used for any further rheological interpretation and for the conversion of measured into rheological data, respectively. Duration or number of data affected by the sphere acceleration regime primarily depends on the specified speed and, to a lesser extent, on the fluid characteristics. The influence of sphere acceleration is almost negligible for very small speeds but becomes very significant for high speeds. For example, for speeds equal to or larger than Ω = 2.25 rps, the sphere acceleration regime covers one third up to half of the sphere path of a BMS experiment.

Fig. 15
figure 15

Basic BMS experiments using the sphere with the diameter D = 12 mm: development of the rotational speed Ω and the torque T measured along the path of one sphere rotation for a specified speed of Ω = 2.25 rps. a Sediment water mixture made of debris flow material with a maximum grain size d max = 1 mm and a sediment concentration C v = 0.502. b Maximum grain size d max = 10 mm and a sediment concentration C v = 0.595 [C v = V s/(V s + V w) with V s = volume of sediment and V w = volume of the water]

Wake formation and viscous overstream

In low viscous fluids (Newtonian or non-Newtonian, particulated or nonparticulated), it was observed that, for very high speeds of Ω = 2.25–4.5 rps, the accelerating sphere generated a wake in front of the sphere and the sphere holder. For higher viscous fluids, such a wake was not observed. For silicon oils with viscosities larger than 2 Pa·s, the oil was lifted up along the sphere holder for higher rotational speeds. This was interpreted as viscous overstream (Ω ≥ 1.35 rps for the medium viscous oil, η = 2 Pa·s, Ω = 0.135 rps in the case of the highly viscous oil η = 60 Pa·s). The measured torque might be biased due to this effect so that corresponding data should be interpreted with caution.

Sphere–particle interaction, data scattering, and temporary jamming

The flow around the measuring sphere is ideally the flow of a homogeneous one-phase system. However, interactions of larger particles with the measuring sphere create a complex flow situation, in particular when both the suspended particles and the measuring sphere are of the same size (Chhabra 2007). Data fluctuation as seen for the different sphere diameters (see next chapter) are, however, related to the different geometries rather than to sphere–sphere interaction. Therefore, we have not considered the interaction of the moving sphere in the suspension.

Data scattering and temporary jamming

Data fluctuation of both the rotational speed and the torque increases with an increasing maximum grain size d max of the particles for the steady drag flow regime (as shown in Fig. 15). For relative maximum grain sizes d max/D ≤ 0.125, the standard deviation of the speed σΩ was typically less than 1% and, including fluctuation, remained <4%. For the same range of the relative maximum grain size, the standard deviation of the torque σT was typically ~3% and, including fluctuation, remained <10%. A strong increase of the standard deviations, σT and σΩ, is observed for relative maximum grain sizes d max/D > 0.125 and for fluids having d max ≥ 5 mm grain size. The standard deviation of the rotational speed, σΩ, not only depends on the relative maximum grain size d max/D but also strongly depends on the sediment concentration C v of the fluid (Fig. 16). For highly concentrated, large-particulated fluids, the standard deviation of the rotational speed σΩ can thus attain values up to 100%. This behavior is mainly due to increasing friction and local or temporary jamming between the sphere, sphere holder, particles, and the container boundary along the sphere path, causing an increasing variation of the speed (Schatzmann et al. 2003b).

Fig. 16
figure 16

Standard deviation of the rotational speed, σΩ, and the torque, σT, vs sediment concentration C v for sediment–water mixtures with variable maximum grain size d max. Average standard deviation values obtained for Ω = 0.0045, 0.0135, 0.045, 0.135, 0.45, 1.35, 2.25/2.7, and 4.5 rps

Influence of prestructured sample

During the first full rotation, the sphere is dragged through an undisturbed sample fluid, whereas, for the following rotations, the sphere is dragged through a prestructured sample due to the influence of the first rotation, namely, along the sphere path. Depending on the value of the measured torque for the first and the following rotations at a given speed, it is, in principle, possible to determine whether or not the fluid is completely relaxed between the first and the following rotations and whether or not the fluid endured fatigue (Schatzmann 2005).

Considerable side effects

The measuring sphere moves approximately at medium container depth through a sample fluid, which is at rest. Even though the sample fluid is stirred before every experiment, settling of particles may take place within measuring time. Although it is not expected that the settling particles influence the torque measurements, one must be aware of an inverse gradient of fluid concentration over the entire container depth. In fluids affected by particle settling, it is, thus, not clear whether at the sphere depth the rheological properties corresponding to the mean concentration of particles are measured.

The relatively small distances between the sphere and the container boundary (side and bottom), as well as the sphere holder, determine the drag flow and contribute to the total measured torque. The sphere holder is 0.6 mm thick and 3 mm long in the direction of the sphere path, and the immerged length varies between s i = 11 mm for a D = 15-mm sphere and s i = 18 mm for a D = 8-mm sphere. The distance between the container side and the sphere is s w = 17 mm, and the distance between the container bottom and the sphere is s b = 22 mm. While the influence of sphere holder and container boundary can be quantified for each sphere dragging across well-defined nonparticulated Newtonian fluids (Schatzmann 2005), some unpredictable effects remain when measuring in particulated Newtonian or non-Newtonian fluids.

One effect is, namely, the temporary transport of larger particles on top of the sphere while being supported backwards by the holder. In this case, the dragging body is not only the sphere and the holder but the sphere, the holder, and the large particle. As a consequence, increased torque values are measured in this case. A second effect is the temporary jamming of particles between the sphere, the container boundary, and/or the holder in high to highly concentrated mixtures producing relatively larger torque values. Both effects are difficult to assess based on the datasets of the torque and the speed in the steady drag flow regime because the effects are usually hidden behind the large data fluctuation in large-particulated fluids. One possibility to detect such an effect is the comparison of the torque data of two or several independent experiments performed at a specified speed. If the torque data of both experiments cover the same range, no such boundary effect is assumed to occur.

Another uncertainty is the influence of the container boundary on the yield stress fluids investigated in this study. The flow field of a sphere dragged across a yield stress fluid is characterized by a sheared zone around the dragged sphere and a nonsheared zone beyond the sheared zone (Beris et al. 1985; Chhabra and Uhlherr 1988). Generally, size and shape of the sheared zone depends on the sphere velocity, the parameters of the rheological model function of the fluid, and the confining boundary. Atapattu et al. (1995) measured size and shape of the sheared zone for spheres dragging across different yield stress fluids in tubes. Beaulne and Mitsoulis (1997) made numerical simulations and found a good agreement between the experimental data of Atapattu et al. (1995) and the numerical results. Based on the results of Beaulne and Mitsoulis (1997), the size of the sheared zone was estimated in the case of the present BMS. It was estimated that, for the different fluids and speeds investigated in the present study, the sheared zone must usually have spread to the container wall and bottom. This might not be the case for the small and medium spheres (D = 8, 12 mm) dragged at low velocities across the rather concentrated yield stress fluids. Because of the lack of appropriate technical equipment to measure the boundary of sheared and nonsheared zones in the present BMS device, this aspect was not further analyzed.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Schatzmann, M., Bezzola, G.R., Minor, HE. et al. Rheometry for large-particulated fluids: analysis of the ball measuring system and comparison to debris flow rheometry. Rheol Acta 48, 715–733 (2009). https://doi.org/10.1007/s00397-009-0364-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00397-009-0364-x

Keywords

Navigation