26.10.2012 Views

CHAPTER 3 MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS

CHAPTER 3 MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS

CHAPTER 3 MECHANICS OF SUSPENSION OF SOLIDS IN LIQUIDS

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>CHAPTER</strong> 3<br />

<strong>MECHANICS</strong> <strong>OF</strong><br />

<strong>SUSPENSION</strong> <strong>OF</strong><br />

<strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

3-0 <strong>IN</strong>TRODUCTION<br />

The physical principles of flow of complex mixtures are based on the interaction between<br />

the different phases, which may mix well or move in superimposed layers. In this chapter,<br />

the basic concepts of motion of particles in a carrying fluid will be presented, as well as the<br />

effect of their concentrations and boundaries. In the previous two chapters, we reviewed the<br />

physical properties of solids, single-phase flows, and some aspects of mixtures of both.<br />

Concepts of non-Newtonian mixtures are reviewed so the reader can understand the<br />

principles used to analyze complex homogeneous flows of very fine particles at high volumetric<br />

concentration.<br />

The physics of solid–liquid mixtures have been the subject of many publications, particularly<br />

by chemical and nuclear engineers. In this chapter, an effort is made to focus on<br />

the practical equations that a slurry engineer may use to accomplish his/her tasks. The engineer<br />

may have to use more than one equation when assessing a mixture to make an engineering<br />

judgment.<br />

3-1 DRAG COEFFICIENT AND TERM<strong>IN</strong>AL<br />

VELOCITY <strong>OF</strong> SUSPENDED SPHERES<br />

<strong>IN</strong> A FLUID<br />

One fundamental aspect to the transportation of solids by a liquid is the resistance, called<br />

the drag force, that such solids will exert, and the ability of the liquid to lift such solids,<br />

called the lift force. Both are complex functions of the speed of the flow, the shape of the<br />

solid particles, the degree of turbulence, and the interaction between particles and the<br />

pipe. One approach is to look at a vehicle that we have all come to know—the airplane.<br />

This distraction from the complex world of slurry flows is justifiable.<br />

3-1-1 The Airplane Analogy<br />

When an airplane flies in a horizontal plane, it is subject to the forces of downward gravity,<br />

upward lift, and drag opposite to its flight path. To maintain steady flight, its engines<br />

3.1


3.2 <strong>CHAPTER</strong> THREE<br />

must develop sufficient thrust to overcome drag. The airplane must also fly above its<br />

stalling speed.<br />

The lift and drag are aerodynamic forces (Figure 3-1). They are proportional to the<br />

surface area, the density of air, the inclination of the airplane body with respect to speed,<br />

and the square of the speed. For the airplane wing, these forces are expressed as<br />

L = 0.5 CL�V 2Sw (3-1)<br />

D = 0.5 CD�V 2Sw (3-2)<br />

where<br />

� = density of the fluid<br />

V = cruising speed of airplane<br />

CL = lift coefficient of wing airfoil<br />

CD = drag coefficient of wing airfoil<br />

The aerodynamic drag consists of two components: the profile drag and induced drag.<br />

The induced drag is proportional to the square of the lift. Airfoils are designed to maximize<br />

the lift-to-drag ratio, or to develop the most lift at the least drag penalty:<br />

2 CD = CD0 + kwC L (3-3)<br />

where<br />

CD0 = the profile drag<br />

kw = a function of the shape of the wing (minimum for an elliptical wing and for a wing<br />

flying in ground effect)<br />

The value of the drag and lift coefficients are determined by the shape of the flying ob-<br />

Thrust<br />

Wing lift<br />

Weight<br />

Forces on an aircraft in<br />

steady horizontal flight<br />

Drag<br />

Stabilizer lift<br />

Weight<br />

Thrust<br />

Drag<br />

Forces on a rocket in<br />

vertical flight<br />

FIGURE 3-1 Lift and drag forces on moving objects.<br />

Buoyancy<br />

Drag<br />

Weight<br />

Forces on a free-falling<br />

particle immersed in a fluid


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

ject, but also by the physical properties of a fluid, particularly the density, viscosity, and<br />

speed of motion. Nondimensional analysis, an important branch of fluid dynamics, allows<br />

the expression of these relationships by characteristic numbers. The Reynolds Number<br />

has already introduced in Chapter 2.<br />

For an airplane in a steady horizontal linear flight, the lift must overcome weight and<br />

the thrust drag. A rocket flying in a vertical plane must develop sufficient thrust to overcome<br />

drag forces as well as weight:<br />

L = W and T = D For an Airplane<br />

T = W + D For a rocket in vertical flight<br />

3-1-2 Buoyancy of Floating Objects<br />

The principle of Archimedes is well known. It states that the buoyancy force developed<br />

by an object static in a fluid is equal to the weight of liquid of equivalent volume occupied<br />

by the object. When the density of the object is less than the density of the liquid, the object<br />

floats, and in the inverse situation, the object sinks.<br />

For a sphere immersed in a fluid of density �L, the buoyancy force is calculated from<br />

the weight of fluid the particle displaces:<br />

3 FBF = (�/6)d g�Lg (3-4)<br />

where<br />

FBF = buoyancy force<br />

dg = sphere diameter<br />

g = acceleration due to gravity (9.78–9.81 m/s2 )<br />

3-1-3 Terminal Velocity of Spherical Particles<br />

Although most solids are not spherical in shape, the sphere is the point of reference for<br />

the analysis of irregularly shaped solids.<br />

3-1-3-1 Terminal Velocity of a Sphere Falling in a Vertical Tube<br />

When a sphere is allowed to fall freely in a tube, the buoyancy and the drag forces act vertically<br />

upward, whereas the weight force acts downward. At the terminal or free settling<br />

velocity, in the absence of any centrifugal, electrostatic, or magnetic forces<br />

W = D + FBF (3-5)<br />

� �dg 3�Sg = � �dg 3 2<br />

� �<br />

�d<br />

2 g<br />

�Lg + 0.5 CD�LV t� � (3-6)<br />

�<br />

6<br />

�<br />

6<br />

�<br />

4<br />

The drag coefficient corresponding to free fall of the particle is calculated as<br />

4(�S – �L)gdg CD = ��<br />

(3-7)<br />

3�LV t 2<br />

where<br />

d g = sphere diameter<br />

g = acceleration due to gravity, typically 9.8 m/s 2 or 32.2 ft/sec 2<br />

3.3


3.4 <strong>CHAPTER</strong> THREE<br />

Vt = the terminal (or free settling) speed<br />

�s = the density of the solid sphere in kg/m3 or slugs/ft3 �L = the density of the liquid<br />

The terminal (or sinking) velocity is measured using a visual accumulation tube with a<br />

recording drum. Various mathematical models have been derived for the drag coefficient.<br />

Turton and Levenspiel (1986) proposed the following equation:<br />

0.413<br />

0.657 CD = (1 + 0.173Re p ) ���<br />

(3-8)<br />

1 + 1.163 × 104 24<br />

� –1.09<br />

Rep<br />

Re p<br />

Example 3-1<br />

Using the Turton and Levenspiel equation, write a small computer program in quickbasic<br />

to tabulate the drag coefficient of a sphere.<br />

LPR<strong>IN</strong>T “ Drag coefficient vs. Reynolds Number based on<br />

Turton, R., and O. Levenspiel”<br />

RE0= 1<br />

15 FOR I=1 TO 10<br />

RE=I*RE0<br />

CD= (24/RE) * (1+0.173*RE^0.657)*(0.413/(1+11630*RE^-1.09)<br />

PR<strong>IN</strong>T US<strong>IN</strong>G “RE= ###### ; Cd = ##.#### “; RE,CD<br />

NEXT I<br />

IF RE>1E6 THEN GOTO 30<br />

RE0=RE<br />

TABLE 3-1 Particle Reynolds Number and Corresponding Drag Coefficient for a<br />

Sphere Based on the Equation of Turton and Levenspiel (1986) as per Example 3-1<br />

Particle Drag Particle Drag Particle Drag<br />

Reynolds coefficient, Reynolds coefficient, Reynolds coefficient,<br />

number, Rep CD number, Rep CD number, Rep CD 1 28.1520 80 1.2266 6000 0.3983<br />

2 15.2735 90 1.1571 7000 0.4042<br />

3 10.8485 100 1.0994 8,000 0.4151<br />

4 8.5809 200 0.5025 9,000 0.4151<br />

5 7.1908 300 0.6793 10,000 0.4200<br />

6 6.2459 400 0.6085 20,000 0.4497<br />

7 5.5588 500 0.5617 30,000 0.4617<br />

8 5.0349 600 0.5281 40,000 0.4671<br />

9 4.6211 700 0.5029 50,000 0.4697<br />

10 4.2851 800 0.4832 60,000 0.4709<br />

20 2.6866 900 0.4675 70,000 0.4713<br />

30 2.0940 1,000 0.4547 80,000 0.4713<br />

40 1.7729 2,000 0.3990 90,000 0.4711<br />

50 1.5670 3,000 0.3878 100,000 0.4707<br />

60 1.4216 4,000 0.3883 200,000 0.4653<br />

70 1.3124 5,000 0.3927 300,000 0.4609


Drag Coefficient C D<br />

25<br />

20<br />

15<br />

10<br />

5<br />

GOTO 15<br />

30 END<br />

0<br />

0 2 4 6 8 10<br />

Rep<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

30 0<br />

C D<br />

C D<br />

0<br />

6<br />

4<br />

2<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

20<br />

200<br />

40<br />

400<br />

Results are tabulated in Table 3-1 and presented in Figure 3-2 in a linear scale rather<br />

than a logarithmic scale. Linear scales are sometimes more useful to the mine operator<br />

who is in a remote area and has little time to waste on difficult logarithmic graphs<br />

3-1-3-2 Very Fine Spheres<br />

For small particles in the range of a diameter d50 < 0.15 mm (0.0059 in), the most common<br />

equation was created by Stokes and reported by Herbich (1991) and Wasp et al. (1977),<br />

who indicate that the main forces are due to the viscosity effect in the laminar flow regime:<br />

D = 3��dg (3-9)<br />

In the laminar regime, the drag coefficient is inversely proportional to the Reynolds number,<br />

i.e., CD = 24/Rep. The terminal velocity is expressed by Stoke’s equation:<br />

2 (�S – �L)d gg Vt = ��<br />

(3-10)<br />

18�L�<br />

Stokes’s equation is limited to particle Reynolds numbers smaller than 0.1, but has often<br />

been used for particle Reynolds Numbers as large as 1 (based on sphere diameter d g).<br />

60<br />

600<br />

80<br />

800<br />

100<br />

Rep<br />

3<br />

10<br />

Rep<br />

CD CD D C<br />

FIGURE 3-2 Drag coefficient of a sphere for Reynolds number smaller than 300,000.<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.6<br />

0.4<br />

0.2<br />

1X10 5<br />

2000<br />

4<br />

2X10<br />

4000<br />

4<br />

4X10<br />

Rep<br />

5<br />

3X10<br />

6000<br />

4<br />

6X10<br />

8000<br />

4<br />

8X10<br />

3.5<br />

10 4<br />

Rep<br />

5<br />

1X10<br />

Rep


3.6 <strong>CHAPTER</strong> THREE<br />

From Equation 3-10, Herbich (1968) pointed out that the radius of particles for which the<br />

validity of the equation is in doubt is expressed as<br />

4.5� 2 � L<br />

� (�S – � L)<br />

R = � � 3/2<br />

This equation is not set in stone for all situations. Rubey (1933) demonstrated one example<br />

by showing that Stoke’s law does not apply to spherical quartz suspended in water<br />

when the particle diameter exceeds 0.014 mm (0.00055 in, mesh 105).<br />

3-1-3-3 Intermediate Spheres<br />

For the range of particle Reynolds numbers between 1 and 1000, i.e., when<br />

dpV0 �<br />

1 < � < 1000<br />

�<br />

Govier and Aziz (1972) reported that Allen (1900) derived the following equation:<br />

(� � – � L)g<br />

Vt = 0.2� � 0.72<br />

��<br />

�L<br />

(3-11)<br />

Example 3-2<br />

A slurry mixture consists of fine rocks at an average particle diameter of 140 �m, with a<br />

particle density of 2800 kg/m3 . The carrier liquid is water with a dynamic viscosity of 1.5<br />

× 10 –3 Pa · s. The volumetric concentration of the solids is 12%. Determine the terminal<br />

velocity of the particles.<br />

Solution<br />

Using Equation 1-9, the dynamic viscosity of the mixture is<br />

�m = �L[1 + 2.5C� + 10.05C� 2 + 0.00273 exp(16.6C�)]<br />

= 1.5 × 10 –3 [1 + 2.5 × 0.12 + 10.05(0.12) 2 + 0.00273 exp (16.6 × 0.12)]<br />

�m = 2.197 × 10 –3 Pa · s.<br />

Let us check the magnitude of the Reynolds number:<br />

= = 4.468<br />

The Allen law applies in a transition regime:<br />

Vt = 0.2 [9.81 × 1.8] 0.72<br />

(140 × 10 –6 ) 1.18<br />

���<br />

(2.197 × 10 –3 /2800) 0.45<br />

140 × 10 –6 × 0.02504 × 2800<br />

���<br />

2.197 × 10 –3<br />

d�V0� �<br />

�<br />

2.83 × 10<br />

Vt = 0.2 × 7.903<br />

–5<br />

��<br />

0.001789<br />

V t = 0.02504 m/s<br />

d p 1.18<br />

� (�/�) 0.45<br />

Richards (1908) demonstrated that Stokes’s equation is inaccurate for particles with a<br />

diameter larger than 0.2 mm (0.00787 in, mesh 70) and conducted extensive tests for


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

quartz particles (with a specific gravity of 2.65) in laminar, transitional, and turbulent<br />

regimes. He derived the following equation for terminal velocity in mm/s:<br />

8.925<br />

Vt = ����<br />

(3-12)<br />

3 1/2 dg{[1 + 95(�S/�L – 1)d g] – 1}<br />

Where dg, the diameter of the sphere, is expressed in mm. This equation covers the range<br />

of particles between 0.15–1.5 mm (0.0059–0.059 in) at particle Reynolds numbers between<br />

10 and 1000.<br />

3-1-3-4 Large Spheres<br />

For particles with a diameter in excess of 1.5 mm, Herbrich (1991) expressed the terminal<br />

velocity by the following equation:<br />

Vt = Kt�[d� ��� g( ��� S/ L�–� 1�)] � (3-13)<br />

where Kt = an experimental constant = 5.45 for Rep > 800, according to Govier and Aziz<br />

(1972).<br />

Equation 3-13 is often called Newton’s law. In the regime of Newton’s law, the drag<br />

coefficient of a sphere is approximately 0.44, as shown in Figure 3-2. Newton’s law applies<br />

to turbulent flow regimes.<br />

Other equations for terminal velocity of particles have been developed by various authors.<br />

Four different equations are presented in Table 3-2.<br />

Example 3-3<br />

Using the Budyruck equation from Table 3-3, determine the terminal velocity of spheres<br />

from 0.1 to1 mm.<br />

A simple computer program is written in quickbasic as follows:<br />

LPR<strong>IN</strong>T<br />

LPR<strong>IN</strong>T “BUDRYCK AND RITT<strong>IN</strong>GER EQUATION FOR TERM<strong>IN</strong>AL<br />

VELOCITY <strong>OF</strong> SPHERES <strong>IN</strong> WATER”<br />

LPR<strong>IN</strong>T<br />

LPR<strong>IN</strong>T DP0 = .1<br />

FOR I=1 to 11<br />

DP = I*DP0<br />

VS= (8.925/DP)*(SQR(1+157*DP^30-1)<br />

LPR<strong>IN</strong>T US<strong>IN</strong>G “PARTICLE DIAMETER = ##.### mm TERM<strong>IN</strong>AL<br />

VELOCITY Vs = ##.### mm/s”;DP,VS<br />

NEXT I<br />

FOR J=12 TO 20<br />

DP = J*DP0<br />

TABLE 3-2 Equations for Terminal Speed of Large Spheres<br />

Name Equation* Application<br />

Budryck 3 1/2 Vt = 8.925[(1 + 157d g) – 1]/dg For dg < 1.1 mm<br />

Rittinger Vt = 87(1.65dg) 1/2 For 1.2 < dg < 2 mm<br />

*Where V t is expressed in mm/s and d g in mm.<br />

3.7


3.8 <strong>CHAPTER</strong> THREE<br />

TABLE 3-3 Calculation of Terminal Velocity of Spheres in Accordance with<br />

Budryck’s Equation<br />

Particle diameter Terminal velocity Particle diameter Terminal velocity<br />

dp in mm Vs in mm/s dp in mm Vs in mm/s<br />

0.1 6.75 0.7 81.63<br />

0.2 22.4 0.8 89.49<br />

0.3 38.34 0.9 96.64<br />

0.4 51.85 1.0 103.26<br />

0.5 63.21 1.1 109.45<br />

0.6 73.02<br />

VS= 87*SQR(1.65*DP)<br />

LPR<strong>IN</strong>T US<strong>IN</strong>G “PARTICLE DIAMETER = ##.### mm TERM<strong>IN</strong>AL<br />

VELOCITY Vs = ##.### mm/s”;DP,VS<br />

NEXT J<br />

END<br />

The results are shown in Tables 3-3, 3-4, and Figure 3-3<br />

Herbich (1968) measured drag coefficients for ocean nodules to be as high as 0.6 at<br />

particle Reynolds numbers of 200. This high value is reached with spheres at a particle<br />

Reynolds number of 1000.<br />

3-1-4 Effects of Cylindrical Walls on Terminal Velocity<br />

The previous paragraphs focused on the settling velocity of a single particle or widely<br />

separated particles. The presence of a vessel or cylindrical walls tends to multiply the interaction<br />

between particles and cause some collisions. Extensive tests have been conducted<br />

on flows in vertical tubes. Brown and associates (1950) recommended multiplying the<br />

terminal speed of a single particle by a wall correction factor Fw. For laminar flows they<br />

proposed to use the Francis equation:<br />

Fw = 1 – (d�/Di) 9/4 (3-14a)<br />

They proposed to use the Munroe equation for a turbulent flow regime:<br />

Fw = 1 – (d�/Di) 1.5 (3-14b)<br />

where Di = the inner diameter of the tube<br />

TABLE 3-4 Calculation of Terminal Velocity of Spheres in Accordance with<br />

Rittinger’s Equation<br />

Particle diameter Terminal velocity Particle diameter Terminal velocity<br />

d p in mm V t in mm/s d p in mm V t in mm/s<br />

1.1 117.21 1.6 141.36<br />

1.2 122.42 1.7 145.71<br />

1.3 127.42 1.8 149.93<br />

1.4 132.23 1.9 154.04<br />

1.5 136.87 2.0 158.04


in mm/s<br />

Terminal velocity V<br />

t<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

in mm/s<br />

Terminal velocity V<br />

t<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 0.01 0.02 0.03 0.04 0.05<br />

Sphere diameter d p in inches<br />

0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Sphere diameter dp<br />

in mm<br />

0.04 0.05 0.06 0.07 0.08<br />

1.0 1.2 1.4 1.6 1.8 2.0<br />

Sphere diameter dp<br />

in mm<br />

(a)<br />

Sphere diameter d p in inches<br />

(b)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Terminal velocity V<br />

t<br />

in inch/sec<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

3.9<br />

inch /sec<br />

Terminal velocity V<br />

t<br />

in<br />

FIGURE 3-3 Terminal velocity of spheres (a) in accordance with Budryck’s equation, (b) in<br />

accordance with Rittinger’s equation.


3.10 <strong>CHAPTER</strong> THREE<br />

Example 3-4<br />

The flow described in Example 3-2 occurs in a 63 mm ID pipe. Determine the corrected<br />

terminal velocity due to the wall effects.<br />

Solution<br />

The terminal velocity was determined to be 0.02504 m/s. The flow is in transition. Equation<br />

3-14a for laminar flow is<br />

Fw = 1 – (d�/DI) 9/4<br />

F w = 1 – (0.140/63) 9/4<br />

Fw = 0.999<br />

Equation 3-14b for turbulent flow is<br />

Fw = 1 – (0.14/63) 1.5 = 0.999.<br />

More recently, Prokunin (1998) extended the analysis of the interaction of the wall<br />

with the motion of a single particle by considering the angle of inclination and any rotation<br />

that the particle may incur. His investigation included immersion in non-Newtonian<br />

flows by testing with glycerin and silicone. He noticed from his tests that when the particle<br />

approaches the wall, it develops a lift force. The lift force seems to increase with a reduction<br />

of the gap that separates the particle from the wall. However, Prokunin could not<br />

explain this lift force and recommended further research.<br />

3-1-5 Effects of the Volumetric Concentration on the<br />

Terminal Velocity<br />

As the volumetric concentration of particles increases, it causes interactions and collisions,<br />

and transfers momentum between the different (finer and coarser) units. The distance<br />

between particles decreases. For spheres at 1% concentration by volume, the interparticle<br />

distance is only 4 diameters. It shrinks to 2.5 diameters at 5% and to 2 diameters<br />

at 10% concentration by volume. In an ideal laminar flow, the interaction is much simpler<br />

than in a turbulent flow.<br />

Worster and Denny (1955) published data on the terminal velocity of coal and gravel<br />

particles, as shown in Table 3-5. The effect of the concentration is clearly marked by a<br />

difference in terminal velocity between a single particle and a volumetric concentration of<br />

30%.<br />

Kearsey and Gill (1963) applied the Carman–Kozeney equation of flow through a<br />

porous medium to determine the terminal velocity as<br />

TABLE 3-5 Terminal Velocity for Coal and Gravel after Worster and Denny (1955)<br />

Coal with a specific gravity of 1.5<br />

________________________________<br />

Gravel with a specific gravity of 2.67<br />

________________________________<br />

Particle size<br />

____________<br />

Single particle 30% Concentration<br />

______________ ________________<br />

Single particle<br />

______________<br />

30% Concentration<br />

________________<br />

mm Inches (cm/s) (ft/s) (cm/s) (ft/s) (cm/s) (ft/s) (cm/s) (ft/s)<br />

1.59 1/16 4.6 0.15 3.0 0.10 9.1 0.30 3.0 0.10<br />

6.4<br />

12.7<br />

1 –4<br />

1 –2<br />

15.2<br />

30.5<br />

1.50<br />

1.00<br />

10.7<br />

21.3<br />

0.35<br />

0.70<br />

30.5<br />

61.0<br />

1.00<br />

2.00<br />

10.7<br />

21.3<br />

0.35<br />

0.70<br />

25.4 1 51.8 1.70 36.6 1.20 106.7 3.50 36.6 1.20


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

(1 – Cv) 1 �P<br />

� � 2 �s p L<br />

where<br />

sp = the specific surface expressed for as sphere as the surface area to volume ratio:<br />

3<br />

�2 KzC v<br />

V c = � �� �� � (3-15)<br />

�d g 2<br />

3.11<br />

sp = = 6/dg Kz = the Kozney constant, which is a function of particle shape, porosity, particle orientation,<br />

and size distribution. The magnitude of Kz is between 3 and 6, but is<br />

commonly assumed to be 5<br />

�P/Li = the pressure gradient in the pipe due to the flow of the mixture<br />

In the process of sedimentation, the pressure gradient is essentially due to the volumetric<br />

concentration of the particles and is expressed as<br />

= Cv(�s – �L)g (3-16)<br />

In addition, the settling velocity due to a volumetric concentration is expressed as<br />

Vc = � �� � (3-17)<br />

For spheres with sp = 6/dg, the equation reduces to<br />

Vc = � �� � (3-18)<br />

As the volumetric concentration increases from 3% to 30%, the velocity drops drastically.<br />

Assuming Kz to be equal to 5.0, the settling velocity for spheres reduces to a simple<br />

equation:<br />

(1 – Cv) = (3-19)<br />

where V0 = the terminal velocity at very low volumetric concentration<br />

Equation 3-19 does not apply to volumetric concentrations smaller than 8%. Equation<br />

3-18 would apply to smaller concentrations.<br />

3<br />

(1 – Cv) (�s – �L) �<br />

�<br />

Vc � �<br />

V0 10Cv<br />

3 (1 – Cv) (�s – �L) �2 �s p<br />

2 gd g<br />

��<br />

36KzCv 3 � 3 (�d g/6) �P<br />

�<br />

Li<br />

g<br />

��<br />

KzCv Example 3-5<br />

Assuming that the terminal velocity at a volumetric concentration of 8% is 100 mm/s,<br />

apply Equation 3-18 from a volumetric concentration of 8–30%. Plot the results in Figure<br />

3-4.<br />

Thomas (1963) proposed the following empirical equation in the range of Vc/V0 of<br />

0.08–1.0:<br />

2.303 log10(Vc/V0) = –5.9CV (3-20)<br />

Example 3-6<br />

The free settling speed of solid particles is 22 mm/s at a volumetric concentration of 1%.<br />

Using the Thomas equation 3-20, determine the settling speed at 25% volumetric concentration.


3.12 <strong>CHAPTER</strong> THREE<br />

V c/<br />

Vo<br />

Solution<br />

2.303 log 10(V c/V 0) = -5.9 × 0.25<br />

V c/V 0 = 10 –0.64<br />

V c/V 0 = 0.2288<br />

V c = 0.2288 × 22 mm/s = 5.03 mm/s<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.1<br />

0.2<br />

Volumetric concentration<br />

FIGURE 3-4 Effect of the volumetric concentration on the terminal velocity of spheres in<br />

accordance with Equation 3-18.<br />

The Kozney-based approach is limited to concentrations where the particles come into<br />

contact with each other in a vertical flow. Beyond this point, the pressure gradient is<br />

smaller than expressed by Equation 3-16. In the case of hard spheres, the settling process<br />

completes when the particles come into contact with each other. In the case of flocculated<br />

particles or clusters of flocculated fluid, stress may cause deformation and further settling<br />

may occur by compaction.<br />

Irregularly shaped particles and flocculates cause the development of a structure with<br />

its own yield stress level. As the particles move closer, the yield stress increases until<br />

equilibrium is reached. The weight of the overburden is then supported by the saturated<br />

fluid and the compacted sediment.<br />

3-2 GENERALIZED DRAG COEFFICIENT—<br />

THE CONCEPT <strong>OF</strong> SHAPE FACTOR<br />

Every day the slurry engineer has to deal with particles of all shapes and sizes. Although<br />

the sphere represents a shape for reference, it is in the minority in the world of crushed or<br />

naturally worn rocks.<br />

Albertson (1953) conducted an extensive study on the effect of the shape of gravel<br />

particles on the fall velocity in a vertical flow (Figure 3-5). He proposed a definition for a<br />

shape factor:<br />

where<br />

a = the longest of three mutually perpendicular axes<br />

b = the third axis<br />

c = the shortest of three mutually perpendicular axes<br />

0.3<br />

c<br />

�A = � (3-21)<br />

�(a�b�)�


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

b<br />

a<br />

direction of fall<br />

FIGURE 3-5 The axes of an irregularly shaped particle, according to Albertson.<br />

3.13<br />

Particles in a free fall tend to align themselves to expose the largest surface to the<br />

flow. In other words, they act as free-falling leaves from a tree on an autumn day, where c<br />

is taken as the dimension opposite to the direction of the fall. The projected area of the<br />

particle is a function of the dimensions “a” and “b” but is often not equaled to such a<br />

product as (ab) because particles are usually not rectangular in shape (see Table 3-6).<br />

In a different approach, Clift et al. (1978) decided to compare the projected area of a<br />

free-falling, irregularly shaped particle, with a sphere of equal projected area in order to<br />

define a diameter:<br />

da = �(4�S� ���)� f /<br />

(3-22)<br />

where<br />

Sf = the projected area of the free-falling particle<br />

However, Albertson (1953) preferred to define a different diameter base, dp, on the<br />

fact that the actual volume of the free-falling particle could be equated to a sphere of the<br />

TABLE 3-6 Clift Shape Factor of Various Particles<br />

Isometric Typical mineral particles<br />

____________________________________ _______________________________________<br />

Particle � c Particle � c<br />

Sphere 0.524 Sand 0.26<br />

Cube 0.694 Sillimanite 0.23<br />

Tetrahedron 0.328 Bituminous Coal 0.23<br />

Irregular Rounded 0.54 Blast Furnace Slag 0.19<br />

Cubic angular 0.47 Limestone 0.16<br />

Tetrahedral 0.38 Talc 0.16<br />

Plumbago 0.16<br />

Gypsum 0.13<br />

Flake Graphite 0.023<br />

Mica 0.003<br />

From Wilson et al. (1992).<br />

c


3.14 <strong>CHAPTER</strong> THREE<br />

same volume but with a diameter of dn. Albertson (1953) therefore proposed a Reynolds<br />

number based on dn: dn�Vt Ren = � (3-23)<br />

�<br />

There may be a marked difference between naturally worn gravel and crushed gravel.<br />

This is a fact that a slurry engineer should bear in mind when extrapolating data from lab<br />

results.<br />

Because Clift chose an equivalent diameter d a based on the projected area, he proposed<br />

a different shape factor:<br />

� c = particle volume/d a 3 (3-24)<br />

Typical values are shown in Table 3-6. The Albertson and Clift shape factors are about<br />

40 years apart in definition but can be related by a factor E:<br />

� c = E� A<br />

(3-25)<br />

The logarithmic curves as shown in Figure 3-6 are sometimes difficult to read. Table<br />

3-7 presents values of drag coefficient versus Reynolds number rounded off to the first<br />

decimal point.<br />

The work of Albertson was developed further by the Inter-Agency Committee on Water<br />

Resources (1958), who developed the following two non-dimensional coefficients<br />

(Figure 3-7):<br />

and<br />

Drag coefficient C D<br />

10.0<br />

1.0<br />

0.1<br />

C N = (� s/� L – 1)g�/V t 3 (3-26a)<br />

C N = 0.75C D/Re n<br />

(3-26b)<br />

C S = �(� s/� L – 1)gd p 3 /(6� 2 ) (3-27a)<br />

C S = 0.125�C DRe n 2 (3-27b)<br />

ALBERTSON SHAPE FACTOR = a/ cb<br />

0.3<br />

0.5<br />

0.7<br />

0 10 100 10 3<br />

10 4<br />

10 5<br />

10 6<br />

0 10 100 103 104 105 106 Particle Reynolds number Re p<br />

FIGURE 3-6 The drag coefficient versus Reynolds number and shape factor. (After Albertson,<br />

1953.)<br />

1.0


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

TABLE 3-7 Drag Coefficient versus Reynolds Number for Different Albertson Shape<br />

Factors<br />

Drag coefficient<br />

3.15<br />

Reynolds number Shape factor = 0.3 Shape factor = 0.5 Shape factor = 0.7 Shape factor = 1.0<br />

7 7.0 6.0 4.7 4.0<br />

8 6.5 5.5 4.3 3.7<br />

9 6.1 5.1 4.0 3.4<br />

10 5.8 4.74 3.75 3.15<br />

15 4.64 3.7 3.0 2.4<br />

20 3.95 3.2 2.55 2.0<br />

32 3.0 2.6 2.1 1.55<br />

40 2.7 2.28 1.84 1.3<br />

50 2.5 2.08 1.67 1.12<br />

60 2.3 1.94 1.56 1.0<br />

70 2.25 1.74 1.4 0.94<br />

80 2.2 1.67 1.35 0.844<br />

100 2.08 1.62 1.3 0.8<br />

150 1.87 1.44 1.16 0.68<br />

200 1.75 1.36 1.11 0.6<br />

300 1.74 1.33 1.08 0.5<br />

400 1.8 1.34 1.09 0.44<br />

500 1.9 1.38 1.1 0.4<br />

600 1.94 1.42 1.12 0.38<br />

700 1.988 1.47 1.14 0.36<br />

800 2.0 1.51 1.15 0.34<br />

900 2.07 1.54 1.16 0.334<br />

1000 2.1 1.58 1.17 0.33<br />

2000 2.3 1.72 1.22 0.3<br />

3000 2.28 1.73 1.19 0.29<br />

4000 2.48 1.69 1.16 0.294<br />

5000 2.21 1.66 1.14 0.3<br />

6000 2.2 1.62 1.13 0.31<br />

7000 2.19 1.58 1.13 0.31<br />

8000 2.183 1.55 1.14 0.32<br />

9000 2.18 1.53 1.14 0.32<br />

The drag coefficient C D is then plotted against the equivalent Reynolds number Re n to<br />

determine the terminal velocity. On a logarithmic scale, C N and C S are superposed as<br />

straight lines for reference (Figure 3-7).<br />

In order to measure the Albertson shape factor, Wasp et al. (1977) developed a correlation<br />

between the sieve diameter and the fall diameter d n (Figure 3-8).<br />

The approach proposed by Albertson and Clift is limited to free fall of particles in a<br />

fluid. However, turbulence can develop new forces. Whenever an engineering contract requires<br />

the drag of particles to be measured, the engineer is well advised to conduct tests in<br />

a fluid of similar dynamic viscosity as the one that will be used in the project. In addition<br />

to the shape factor and drag coefficient, the slurry engineer must also determine the fluid<br />

density, dynamic viscosity at the temperature of pumping, particle density (or specific<br />

gravity of solids), nominal (or statistical average) diameter, and fall velocity.


Sieve diameter (mm)<br />

3.16 <strong>CHAPTER</strong> THREE<br />

FIGURE 3-7 C D and C W versus particle Reynolds number for different shape factors. Adapted<br />

from the Inter-Agency Committee on Water Resources (1958).<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

spheres<br />

S.F = 0.3<br />

S.F = 0.5<br />

S.F = 0.7<br />

S.F= 0.9<br />

0 0.2 0.4 0.6 0.8 1.0<br />

Sieve diameter (mm)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

S.F = 0.3<br />

S.F =0.7<br />

S.F =0.5<br />

spheres<br />

0 1 2 3 4<br />

Fall diameter (mm) Fall diameter (mm)<br />

FIGURE 3-8 Relationship between sieve and fall diameter after Wasp et al. (1977).<br />

S.F=0.9<br />

5


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

Example 3-7<br />

A naturally worn particle has an Albertson shape factor of 0.7. It has a nominal diameter<br />

of 250 �m. Its density is 3000 kg/m3 . It is allowed to free-fall in water at a temperature of<br />

25° C.<br />

Calculate the fall velocity for the single particle and the fall velocity if the volumetric<br />

concentration of particles is increased to 20%.<br />

Solution<br />

Referring to Table 2-7 (or Table 2-8 for USCS units), the kinematic viscosity of water is<br />

0.89 × 10 –6 m2 2 /s. We need to determine the coefficient CS = 0.125�CD/Ren. The curves<br />

published by Inter-Agency Committee on Water Resources indicate that CS =<br />

2 3 2 0.125�CD/Ren = 0.167�(�s/�L – 1)gd p/� = 203.<br />

From Figure 3-6, at a shape factor of 0.7 and CS of 203, the Reynolds number would<br />

be 7.2Vt = Re/(�dp) = 7.2/(890,000 × 0.00025) = 0.0324 m/s for a single particle.<br />

Applying Equation 3-18 for a concentration of 20%, the velocity would be 0.256 ×<br />

0.0324 = 0.0083 m/s.<br />

3-3 NON-NEWTONIAN SLURRIES<br />

3.17<br />

Various models have been developed over the years to classify complex two- and threephase<br />

mixtures (Table 3-8). In the case of mining, the following mixtures are often encountered:<br />

� A fine dispersion containing small particles of a solid, which are uniformly distributed<br />

in a continuous fluid and are found in copper concentrate pipelines and in slurry from<br />

grinding after classification, etc.<br />

TABLE 3-8 Regimes of Flows for Newtonian and Non-Newtonian Mixtures after<br />

Govier and Aziz (1972)<br />

Single-phase flows<br />

___________________________<br />

Multiphase flows (gas–liquid, liquid–liquid,<br />

gas–solid, liquid–liquid)<br />

___________________________________________________<br />

Single-phase behavior<br />

_____________________________________________________<br />

Multiphase behavior<br />

___________________________<br />

Pseudohomogeneous<br />

_______________________________<br />

Heterogeneous<br />

__________________<br />

True homogeneous Laminar, transition, and<br />

turbulent flow regime<br />

Turbulent flow regime only<br />

Purely viscous Newtonian flows<br />

Purely viscous, non-Newtonian, Bingham plastic<br />

and time-independent Dilatant<br />

Pseudoplastic<br />

Yield pseudoplastic<br />

Purely viscous, non-Newtonian Thixotropic<br />

and time-dependent Rheopectic<br />

Viscoelastic Many forms


3.18 <strong>CHAPTER</strong> THREE<br />

� A coarse dispersion containing large particles distributed in a continuous fluid and encountered<br />

in SAG mills, cyclone underflows, and in certain tailings lines, etc.<br />

� A macro-mixed flow pattern containing either a frothy or highly turbulent mixture of<br />

gas and liquid, or two immiscible liquids under conditions in which neither is continuous.<br />

Such patterns are found in flotation circuits in which froth is used to separate concentrate<br />

from gangue.<br />

� A stratified flow pattern containing a gas, liquid, two slurries of different particle sizes,<br />

or two immiscible liquids under conditions in which both phases are continuous.<br />

Designing a pipeline to operate in a non-Newtonian flow regime must be based on reliable<br />

test data about the rheology and particle sizing (see Table 3-9). The engineer must<br />

be cautious before venturing into generalizations about rheological properties.<br />

In Figure 1-4 of Chapter 1, the relationship between dynamic viscosity and volumetric<br />

concentration was presented. In fact, the industry has accepted the criterion that friction<br />

losses are highly dependent on slurry viscosity in cases where the average particle diameter<br />

is finer than 40–60 microns, and (depending on the specific gravity) at volumetric concentrations<br />

in excess of 30%.<br />

Fibrous slurries such as fermentation broths, fruit pulps, crushed meal animal feed,<br />

tomato puree, sewage sludge, and paper pulp may not contain a high percentage of solids,<br />

but may flow as non-Newtonian regimes. With these materials, the long fibers are flexible<br />

and intertwine into a close-packed configuration and entrap the suspending medium. The<br />

fibers may be flocculated or may form flocs with an open structure. Based on the volume<br />

content of the flocs, the mixture may develop high dynamic viscosity. However, because<br />

the flocs are compressible, they may deform with the flow.<br />

Flocculated slurries are encountered in flotation cells circuits, thickeners, and various<br />

processes in mineral extraction plants. With the formation of flocs, the slurry may develop<br />

an internal structure. This structure may develop properties leading to a non-Newtonian<br />

flow, shear thinning behavior (pseudoplastic), and sometimes thixotropic time-dependent<br />

behavior. When shear stresses are applied to the slurry, the floc sizes may shrink and become<br />

less capable of entrapping the carrier slurry. At higher shear stresses, the flocs may<br />

shrink to the size of particles, and the flow may lose its non-Newtonian behavior.<br />

3-4 TIME-<strong>IN</strong>DEPENDENT NON-NEWTONIAN<br />

MIXTURES<br />

Certain slurries require a minimum level of stress before they can flow. An example is<br />

fresh concrete that does not flow unless the angle of the chute exceeds a certain minimum.<br />

Such a mixture is said to posses a yield stress magnitude that must be exceeded before<br />

that flow can commence. A number of flows such as Bingham plastics, pseudoplastics,<br />

yield pseudoplastics, and dilatant are classified as time-independent non-Newtonian fluids.<br />

The relationship of wall shear stress versus shear rate is of the type shown in Figure<br />

3-9 (a), and the relationship between the apparent viscosity and the shear rate is shown in<br />

Figure 3-9 (b). The apparent viscosity is defined as<br />

�a = Cw/(d�/dt) (3.28)<br />

3-4-1 Bingham Plastics<br />

For a Bingham plastics it is essential to overcome a yield stress � 0 before the fluid is set in<br />

motion. The shear stress versus shear rate is then expressed as


TABLE 3-9 Examples of Bingham Slurries<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

Coefficient<br />

Yield of rigidity,<br />

Particle size, Density, Stress, � mPa · s<br />

Slurry d 50 kg/m 3 Pa (cP) Reference<br />

3.19<br />

54.3% Aqueous suspension 92% under 74 �m 1520 3.8 6.86 Hedstrom (1952)<br />

of cement, rock<br />

Flocculated aqueous China 80% under 1 �m 1280 59 13.1 Valentik &<br />

clay suspension No. 1 Whitmore (1965)<br />

Flocculated aqueous China 80% under 1 �m 1207 25 6.7 Valentik &<br />

clay suspension No. 4 Whitmore (1965)<br />

Flocculated aqueous China 80% under 1 �m 1149 7.8 4.0 Valentik &<br />

clay suspension No. 6 Whitmore (1965)<br />

Aqueous clay suspension I 1520 34.5 44.7 Caldwell &<br />

Babitt (1941)<br />

Aqueous clay suspension III 1440 20 32.8 Caldwell &<br />

Babitt (1941)<br />

Aqueous clay suspension V 1360 6.65 19.4 Caldwell &<br />

Babitt (1941)<br />

Fine coal @ 49% C W 50% under 40 �m 1 5 Wells (1991)<br />

Fine coal @ 68% C W 50% under 40 �m 8.3 40 Wells (1991)<br />

Coal tails @ 31% C W 50% under 70 �m 2 60 Wells (1991)<br />

Copper concentrate @ 50% under 35 �m 19 18 Wells (1991)<br />

48% C W<br />

21.4% Bauxite < 200�m 1163 8.5 4.1 Boger & Nguyen<br />

(1987)<br />

Gold tails @ 31% C W 50% under 50 �m 5 87 Wells (1991)<br />

18% Iron oxide < 50 �m 1170 0.78 4.5 Cheng &<br />

Whittaker (1972)<br />

7.5 % Kaolin clay Colloidal 1103 7.5 5 Thomas (1981)<br />

Kaolin @ 32% C W 50% under 0.8 �m 20 30 Wells (1991)<br />

Kaolin @ 53% CW with 50% under 0.8 �m 6 15 Wells (1991)<br />

sodium silicate<br />

Kimbelite tails @ 37% C W 50% under 15 �m 11.6 6 Wells (1991)<br />

58% Limestone < 160 �m 1530 2.5 15 Cheng &<br />

Whittaker (1972)<br />

52.4% Fine liminite < 50 �m 2435 30 16 Mun (1988)<br />

Mineral sands tails @ 50% under 160 �m 30 250 Wells (1991)<br />

58% C w<br />

13.9% Milicz clay < 70 �m 2.3 8.7 Parzonka (1964)<br />

16.8% Milicz clay < 70 �m 5.3 13.6 Parzonka (1964)<br />

19.6% Milicz clay < 70 �m 13 25 Parzonka (1964)<br />

Phosphate tails @ 37% C W 85% under 10 �m 28.5 14 Wells (1991)<br />

14% Sewage sludge 1060 3.1 24.5 Caldwell &<br />

Babitt (1941)<br />

Red mud @ 39% C W 5% under 150 �m 23 30 Wells (1991)<br />

Zinc concentrate @ 75% C W 50% under 20 �m 12 31 Wells (1991)<br />

Uranium tails @ 58% C W 50% under 38 �m 4 15 Wells (1991)


3.20 <strong>CHAPTER</strong> THREE<br />

Shear Stress �<br />

Apparent viscosity � a<br />

Bingham Plastic<br />

Dilatant<br />

Newtonian<br />

Rate of shear (� = du/dy)<br />

Bingham Plastic<br />

Yield Pseudoplastic<br />

Pseudoplastic<br />

Dilatant<br />

Newtonian<br />

Pseudoplastic<br />

Rate of shear (� = du/dy)<br />

(b)<br />

FIGURE 3-9 (a) Shear stress versus shear rate; (b) viscosity versus shear rate of time-independent<br />

non-Newtonian fluids.


�w – �0 = �d�/dt (3-29)<br />

where<br />

�w = shear stress at the wall<br />

�0 = yield stress<br />

� = the coefficient of rigidity or non-Newtonian viscosity<br />

It is also related to a Bingham plastic limiting viscosity at infinite shear rate by the following<br />

equation:<br />

�0 � = � + �� (3-30)<br />

(d�/dt)<br />

The magnitude of the yield stress �0 may be as low as 0.01 Pascal for sewage sludge<br />

(Dick and Ewing, 1967) or as high as 1000 MPa for asphalts and Bitumen (Pilpel, 1965).<br />

The coefficient of rigidity may be as low as the viscosity of water or as high as 1000 poise<br />

(100 Pa · s) for some paints and much higher for asphalts and bitumen. In the case of tarbased<br />

emulsions or certain tar sands, it is customary to add certain chemicals to reduce the<br />

dynamic viscosity of the emulsion or the coefficient of rigidity of the slurry. Tables 3-9<br />

presents examples of Bingham slurries, magnitudes of yield stress, and coefficients of<br />

rigidity � values.<br />

Example 3-8<br />

Samples of a mineral slurry with C w = 45% are examined in a lab. From the measurements<br />

of the rate of shear (�) and shear stress (�), determine the yield stress and viscosity.<br />

Rate of Shear � [s –1 ] 100 150 200 300 400 500 600 700 800<br />

Shear Stress � (Pa) 10.93 12.27 13.49 15.68 17.66 19.49 21.2 22.84 24.43<br />

� – � 0 (Pa) 4.11 5.45 6.67 8.87 10.85 12.67 14.39 16.03 17.61<br />

The data is plotted in Figure 3-10. At a low shear rate < 100s – 1, the slope is<br />

At high shear rate<br />

� = 4.426/100 = 0.0443 Pa · s<br />

4.426<br />

�� = � = 0.0164 Pa · s<br />

270<br />

� =<br />

Take a point at high shear rate (700 s –1 ):<br />

Check at du/dy = 600<br />

at du/dy = 800<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

� =<br />

� w – � 0<br />

� du/dy<br />

16.03<br />

� 700<br />

� = 0.0229 Pa · s<br />

14.394<br />

� = � = 0.02399<br />

600<br />

3.21


Shear stress (Pa)<br />

3.22 <strong>CHAPTER</strong> THREE<br />

17.61<br />

� = � 0.022<br />

800<br />

An average � = 0.023 Pa · s is taken.<br />

Alternative � = �0/(du/dy) + �a � = 6.82/700 + 0.0164 = 0.026 Pa · s<br />

This example shows that at zero rate of shear the shear stress is 6.82 Pa. The yield<br />

stress is therefore 6.82 Pa.<br />

The yield stress increases as the concentration of solids augments. Thomas (1961) proposed<br />

the following relationships between yield stress �0, coefficient of rigidity �, concentration<br />

by volume Cv, and viscosity of the suspending medium �:<br />

� 0 = K 1C v 3 (3-31)<br />

�/� = exp(K2Cv) (3-32)<br />

where K1 and K2 = constants and are characteristics of the particle size, shape, and concentration<br />

of the electrolyte concentration.<br />

These equations were derived from the work of Thomas (1961) on suspensions of titanium<br />

dioxide, graphite, kaolin, and thorium oxide in a range of particle sizes from<br />

0.35–13 micrometers and in volume concentration of 2–23%.<br />

Thomas (1961) defined a shape factor �T1 for nonspherical particles as<br />

�T1 = exp[0.7(sp/s0 – 1)] (3-33)<br />

where<br />

sp = the surface area per unit volume of the actual particles<br />

s0 = the surface area per unit volume of a sphere of equivalent dimensions or 6/dg He indicated that the coefficient K 1 might then be expressed as<br />

30<br />

28<br />

24<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0<br />

0 100 200 300<br />

400 500 600 700<br />

FIGURE 3-10 Plot of data for Example 3-8.<br />

800 900<br />

Rate of shear (sec<br />

-1<br />

)


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

u�T1 �2 d p<br />

3.23<br />

K 1 = (3-34)<br />

Where K 1 is expressed in Pa (or lb f/ft 2 with u = 210 in), and the particle diameter d p is expressed<br />

in microns.<br />

Thomas defined a second shape factor � T2 = (s p/s 0) 1/2 to derive the equation:<br />

K 2 = 2.5 + 14� T 2/�d� p� when 0.4 < d p < 20 microns (3-35)<br />

Thomas (1963) extended his work to flocculated mixtures with dispersed fine and ultrafine<br />

particles with overall dimensions up to 115 microns. He derived the following equations:<br />

�/� = exp[(2.5 + �)Cv] (3-36)<br />

where<br />

� = �[( �d� f /d� ap� p) � –� 1�]� (3-37)<br />

where<br />

� = the ratio of immobilized dispersing fluid to the volume of solids related approximately<br />

to the particle and floc apparent diameter<br />

df = the apparent floc diameter<br />

dapp = the apparent particle diameter<br />

This particle diameter is shown by the following:<br />

dapp = dp(s0/sp) exp(– 1 –<br />

2 ln2 �) (3-38)<br />

where<br />

� = the logarithmic standard deviation<br />

In general, and at a constant temperature, the following equations are applied to Bingham<br />

plastic slurries:<br />

�/� = A exp(BCv) (3-39)<br />

�0 = E exp(FCv) (3-40)<br />

The constants A, B, E, and F are derived from tests measuring particle size, shape, and the<br />

nature of their surface.<br />

Gay et al. (1969) proposed the following correlation for high concentrations of solids:<br />

�/� = exp{[2.5 + [Cv/(Cv� – Cv)] 0.48 ](Cv/Cv�)} (3-41)<br />

where<br />

Cv� = the maximum packing concentration of solids<br />

For a change in temperature in the order of 27°C (50°F). Parzonka (1964) developed<br />

the following power law equation:<br />

–n � = K3T a (3-42a)<br />

where<br />

n = an exponent<br />

K3 = an exponent<br />

Ta = absolute temperature<br />

Govier and Aziz (1972) proposed an equation based on an exponential drop of Bingham<br />

plastic viscosity with temperature:


3.24 <strong>CHAPTER</strong> THREE<br />

� = A exp(B/T) (3-42b)<br />

To obtain the viscosity, plot the curve of the shear stress (� – �0) in Pascals against the<br />

shear rate � (s –1 ).<br />

3-4-2 Pseudoplastic Slurries<br />

Pseudoplastic fluids are time-independent non-Newtonian fluids that are characterized by<br />

the following:<br />

� An infinitesimal shear stress, which is sufficient to initiate motion<br />

� The rate of increase of shear stress with respect to the velocity gradient decreases as<br />

the velocity gradient increases<br />

This type of flow is encountered when fine particles form loosely bound aggregates<br />

that are aligned, stable, and reproducible at a given magnitude of shear rate.<br />

The behavior of pseudoplastic fluids is difficult to define accurately. Various empirical<br />

equations have been developed over the years and involve at least two empirical factors,<br />

one of which is an exponent. For these reasons, pseudoplastic slurries are often<br />

called power-law slurries. The shear stress is defined in terms of the shear rate by the following<br />

equation:<br />

�w = K[(d�/dt) n ] (3-43)<br />

where<br />

K = the power law consistency factor, expressed in Pa · sn n = the power law behavior index, and is smaller than unity<br />

Examples of pseudoplastic slurries are shown in Table 3-10.<br />

The apparent viscosity of a pseudoplastic is defined in terms of the ratio of the shear<br />

stress to the shear rate:<br />

� a = [� w/(d�/dt)] (3-44)<br />

3-4-2-1 Homogeneous Pseudoplastics<br />

Pseudoplastic slurries are another category of non-Newtonian slurries. Pseudoplastics are<br />

divided into homogeneous and pseudohomogeneous mixtures. Whereas in the case of a<br />

Bingham slurry, it was pointed out that the coefficient of rigidity was a linear function of<br />

the shear rate, in the case of a pseudoplastic, the coefficient of rigidity is expressed by the<br />

following power law:<br />

� = K(d�/dt) n–1 (3-45)<br />

The shear stress is plotted against the shear rate on a logarithmic scale at various volume<br />

fractions. From the slope of such a plot, “K,” the power law consistency factor, and<br />

“n,” the power law behavior index (smaller than unity) are derived as plotted in Figure 3-<br />

11.<br />

As indicated in Figure 3-12 the magnitude “K,” the power law consistency factor, and<br />

the power law factor index n are dependent on the volumetric concentration of solids.<br />

Example 3-9<br />

A phosphate slurry mixture is tested using a rheogram. The following data describe the<br />

relationship between the wall shear stress and the shear rate:


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

d�/dt 0 50 100 150 200 300 400 500 600 700 800<br />

�w(Pa) 25 32 43 51 53 56 58 60 62 63.2 64.3<br />

The mixture is non-Newtonian. If it is considered a power law slurry, derive the power<br />

law exponent “n” and the power law coefficient K.<br />

Solution<br />

The first step is to plot the data on a logarithmic scale. In the equation for a pseudoplastic,<br />

the coefficient of rigidity is expressed by equations (3.43) and (3.45), the values of “K”<br />

and “n.” By using the logarithmic scale:<br />

log �w = log K + n log (d�/dt)<br />

log(d�/dt) 1.699 2 2.176 2.301 2.477 2.602 2.669 2.778 2.845 2.903<br />

log(�w) 1.505 1.633 1.707 1.724 1.748 1.763 1.778 1.792 1.8 1.808<br />

n — 0.425 0.592 0.136 0.136 0.12 0.154 0.112 0.13 0.14<br />

log(d�/dt) 2 – log(d�/dt) 1<br />

n = ���<br />

(log�w) 2 – (log�w) 1<br />

n � 0.132<br />

1.8 = log K = 0.132 × 2.843<br />

log K = 1.424<br />

K = 26.5<br />

TABLE 3-10 Examples of Power Law Pseudoplastics<br />

Range of Range of Angle of<br />

Particle weight consistency flow<br />

size, concentration, coefficient K, behavior<br />

Slurry d 50 % Ns n /m 2 index, n Reference<br />

3.25<br />

Cellulose acetate 1.5–7.4 1.4–34.0 0.38–0.43 Heywood (1996)<br />

Drilling mud—barite 14.7 �m 1.0–40.0 0.8–1.3 0.43–0.62 Heywood (1996)<br />

Sand in drilling mud 180 �m 1.0–15% 0.72–1.21 0.48–0.57 Heywood (1996)<br />

sand using<br />

drilling mud<br />

with 18%<br />

barite<br />

Graphite 16.1 �m 0.5–5.0 Unknown Probably 1 Heywood (1996)<br />

Graphite and 5 �m 32.2 total<br />

magnesium (4.1 graphite 5.22 0.16 Heywood (1996)<br />

hydroxide and 28.1<br />

magnesium<br />

hydroxide)<br />

Flocculated kaolin 0.75 �m 8.9–36.3 0.3–39 0.117–0.285 Heywood (1996)<br />

Deflocculated kaolin 0.75 �m 31.3–63.7 0.011–0.6 0.82–1.56 Heywood (1996)<br />

Magnesium hydroxide 5 �m 8.4–45.3 0.5–68 0.12–0.16 Heywood (1996)<br />

Pulverized fuel ash 38 �m 63–71.8 3.3–9.3 0.44–0.46 Heywood (1996)<br />

(PFA-P)<br />

Pulverized fuel ash 20 �m 70–74.4 2.12–9.02 0.48–0.57 Heywood (1996)<br />

(PFA-P)


3.26 <strong>CHAPTER</strong> THREE<br />

Shear stress<br />

(in units of pressure)<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

0.0001<br />

slope = y/x<br />

Consider d�/dt = 700. Check �w = K(d�/dt) n .<br />

62.9 = 26.5 × 7000.132 This is close to the measured stress of 63.2 Pa. Therefore, the equation of this phosphate<br />

slurry is:<br />

�w = 26.5(d�/dt) 0.132<br />

The coefficient of rigidity is obtained as:<br />

Power Law Consistency Factor K<br />

Pa.s<br />

n<br />

/cm<br />

2<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0<br />

x<br />

0 1 10 100 1000 10,000<br />

Shear rate (1/sec)<br />

20<br />

clays<br />

magnetite<br />

40<br />

Volume Fraction of<br />

solids, C V<br />

y<br />

0<br />

n = y/x<br />

FIGURE 3-11 Plotting the rheology on a logarithmic scale to obtain the consistency factor<br />

“K” and the flow behavior index “n” of Pseudoplastics.<br />

Flow Behavior Index "n"<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0<br />

magnetite<br />

clays<br />

20 40<br />

K<br />

Volume Fraction of<br />

solids, CV<br />

FIGURE 3-12 Effect of volumetric concentration on the consistency factor “K” and the flow<br />

behavior index “n” of Pseudoplastics (after Aziz and Govier, 1972).<br />

n


at d�/dt = 700<br />

at d�/dt = 600.<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

� = K(d�/dt) n–1<br />

� = 26.5(d�/dt) –0.878<br />

� = 26.5 × (700) –0.878<br />

� = 0.084 Pa · s<br />

� = 26.5 × 600 = 0.096 Pa · s<br />

3.27<br />

3-4-2-2 Pseudohomogeneous Pseudoplastics<br />

Pseudohomogeneous pseudoplastics behave similarly to their homogeneous counterparts.<br />

Clay suspensions and magnetite-based slurries demonstrate an exponential relationship<br />

between n and C v as shown in Figure 3-12. The power law factor K has a more complex<br />

relationship with C v, as shown in Figure 3-12.<br />

Various equations have been derived to solve the power law factor of pseudoplastics.<br />

These equations are presented to help the reader appreciate the rheological constants that<br />

must be determined by testing, as will be described in Section 3-6.<br />

The Prandtl–Eyring equation is based on Dahlgreen’s (1958) discussion of the study<br />

conducted by Eyring and Prandtl on the kinetic theory of liquids:<br />

� = A sinh –1 [(d�/dt)/B] (3-46)<br />

where<br />

A and B = the rheological constants<br />

sinh = the hyperbolic function<br />

From Equation 3-44, the apparent viscosity is derived as<br />

�a = {A/(d�/dt)}{sinh –1 [(d�/dt)/B]} (3-47)<br />

The Ellis equation is more flexible but is an empirical equation and uses three rheological<br />

constants. Skelland (1967) demonstrates how the equation is based on the work of Ellis<br />

and Round and is explicit with respect to the velocity gradient rather than the shear rate:<br />

(d�/dt) = (A0 + A1� (�–1) )�w (3-48)<br />

where A0, A1, and � are the rheological coefficients of the slurry material.<br />

The apparent viscosity is expressed as<br />

(�–1) �a = 1/(A0 + A1�w ) (3-49)<br />

When A1 = 0, the equation takes on a Newtonian form where A0 = 1/�.<br />

The equation reduces to the conventional power law equation with � = 1/n and A1 =<br />

(1/k) 1/n . When � > 1, the equation approaches a Newtonian flow at low shear stresses, and<br />

when � < 1, it tends to approach a Newtonian flow at high shear stress.<br />

The Cross equation (Cross, 1965) is a versatile equation that is based on measurements<br />

of viscosity, �0 at zero shear rate and �� at infinite shear rates.<br />

�� – �0 �a = �0 + ��<br />

(3-50)<br />

2/3<br />

1 + �(d�/dt)<br />

where � is a coefficient used to express to the shear stability of the mixture.<br />

This equation has been tested and has successfully predicted the behavior of a wide


3.28 <strong>CHAPTER</strong> THREE<br />

variety of pseudoplastic mixtures, such as suspensions of limestone, non-aqueous polymer<br />

solutions, and nonaqueous pigment paste.<br />

3-4-3 Dilatant Slurries<br />

Dilatant fluids are time-independent non-Newtonian fluids and are characterized by the<br />

following:<br />

� An infinitesimal shear stress is sufficient to initiate motion.<br />

� The rate of increase of shear stress with respect to the velocity gradient increases as the<br />

velocity gradient increases.<br />

Dilatant fluids, therefore, use similar equations as pseudoplastic fluids. They are much<br />

less common than pseudoplastics. Dilatancy is observed under specific conditions such as<br />

certain concentrations of solids, shear rates, and the shape of particles. Dilatancy is due to<br />

the shift, under shear action, of a close packing of particles to a more open distribution in<br />

the liquid.<br />

Govier et al. (1957) observed the phenomena of dilatancy in suspensions of magnetite,<br />

galena, and ferrosilicon in a range of particle sizes from 5 microns to 70 microns.<br />

It is observed that the slope of the shear stress versus the shear rate increases, particularly<br />

in the range of shear rates from 80 to 120 sec –1 . Metzener and Whitlock (1958) explained<br />

the phenomenon of dilatancy as follows.<br />

Two mechanisms account for the inflection and subsequent increase in the slope of<br />

the curve. Initially, the shear stress approaches a magnitude at which the size of flowing<br />

particles and aggregates is at a minimum and a Newtonian behavior develops (at<br />

the inflection of the curve). As the level of stress rises, the mixture expands volumetrically,<br />

and entire layers of particles start to slide or glide over each other. In the interim,<br />

the slurry acts as a pseudoplastic until the shear stress is high enough to cause dilatancy.<br />

The phenomenon of dilatancy is not easy to model. According to Metzener and Whitlock<br />

(1958), it is observed at volumetric concentration in excess of 27–30% and at shear<br />

rates in excess of 100 s –1 .<br />

3-4-4 Yield Pseudoplastic Slurries<br />

Yield pseudoplastic fluids are time-independent non-Newtonian fluids and are characterized<br />

by the following:<br />

� An infinitesimal shear stress is sufficient to initiate motion.<br />

� The rate of increase of shear stress, with respect to the velocity gradient, decreases as<br />

the velocity gradient increases.<br />

� A yield stress must be overcome at zero shear rate for motion to occur.<br />

Examples of yield pseudoplastics are shown in Table 3-11.<br />

Equation 3-44 is then modified to account for the yield stress as follows:<br />

�w – �0 = K[(d�/dt) n ] (3-51)<br />

Equation 3-51 is known as the Herschel–Buckley equation of yield pseudoplastics and<br />

is accepted by most slurry experts to describe the rheology of yield pseudoplastics with


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

TABLE 3-11 Examples of Yield Pseudoplastics<br />

Range of Angle of<br />

consistency flow<br />

Density, Yield stress coefficient K, behavior<br />

Slurry kg/m 3 � 0, Pa Ns n /m 2 index, n Reference<br />

3.29<br />

Sewage sludge 1024 1.268 0.214 0.613 Chilton and Stainsby (1998)<br />

Sewage sludge 1011 0.727 0.069 0.664 Chilton and Stainsby (1998)<br />

Sewage sludge 1013 2.827 0.047 0.806 Chilton and Stainsby (1998)<br />

Sewage sludge 1016 1.273 0.189 0.594 Chilton and Stainsby (1998)<br />

Kaolin slurry 1071 1.880 0.010 0.843 Chilton and Stainsby (1998)<br />

Kaolin slurry 1061 1.040 0.014 0.803 Chilton and Stainsby (1998)<br />

Kaolin slurry 1105 4.180 0.035 0.719 Chilton and Stainsby (1998)<br />

low to moderate concentration of solids. At high shear rates, certain complex phenomena<br />

such as dilatancy may develop. Certain bentonite clays develop a yield pseudoplastic rheology<br />

at 20% concentration by volume.<br />

Krusteva (1998) investigated the rheology of a number of inorganic waste slurries<br />

such as drilling fluids in petroleum output, residue mineral materials in tailing ponds, filling<br />

of abandoned mine galleries, etc. In the case of clay containing industrial wastes, he<br />

indicated that colloidal forces of attraction or repulsion are ever present with Brownian<br />

forces and may cause thermodynamic instability. Waste materials such as blast furnace<br />

slag, fly ash, and material from mine filling exhibited various forms of a yield pseudoplastic<br />

rheology.<br />

The behavior of yield pseudoplastics can be expressed by the Carson model as described<br />

by Lapasin et al. (1998):<br />

�n = � n n<br />

0 +�� (d�/dt) (3-52)<br />

By binary system, Lapasin meant a mixture of two sizes of particles above the colloidal<br />

range and by ternary, three sizes. Alumina powders with average d50 diameters of 0.9<br />

�m, 1.4 �m, and 3.9 �m, and different specific surface areas (8.23 m2 /cm3 , 5.74<br />

m2 /cm3 , and 2.65 m2 /cm3 ) were investigated. A dispersing agent was used. Appreciable<br />

time-dependent effects were only noticed at a concentration of the dispersing agent below<br />

a critical value. Multicomponent suspensions were found to have a viscosity that<br />

was dependent on the total volume concentration of solids Cv and on the composition of<br />

the dispersed phase expressed as a volume fraction. It was also dependent on the shear<br />

rate of the mixture.<br />

Vlasak et al. (1998) investigated the addition of peptizing agents to kaolin–water mixtures.<br />

These mixtures were described as yield pseudoplastics that follow the<br />

Bulkley–Herschel rheological model (these will be discussed in Chapter 5). The addition<br />

of peptizing agents initially achieved a rapid drop of viscosity down to 8–10% of the original<br />

value up to an optimum concentration. As the concentration of the peptizing agent is<br />

increased beyond an optimum value, its effects are neutralized and the viscosity of the<br />

slurry increases again. Soda Water-GlassTM as a peptizing agent seemed to achieve the<br />

best reduction in viscosity when added at a concentration of 0.4%. The effect was a drastic<br />

drop of viscosity by 92% of its original value (without the peptizing agent). The optimum<br />

concentration of sodium carbonate, another peptizing agent, was 0.1%. The viscosity<br />

was reduced by 90%. These narrow bands of concentration of peptizing agents can<br />

effectively reduce the cost of hydro-transporting kaolin–water mixtures by reducing viscosity<br />

and therefore the coefficient of friction.


3.30 <strong>CHAPTER</strong> THREE<br />

3-5 TIME-DEPENDENT NON-NEWTONIAN<br />

MIXTURES<br />

Because crude oils and slurries of tar sands from certain Canadian mining projects develop<br />

a time-dependent non-Newtonian behavior in cold temperatures, a section of this chapter<br />

will pay attention to these complex thixotropic properties.<br />

In time-dependent non-Newtonian flows, the structure of the mixture and the orientation<br />

of particles are sensitive to the shear rates. Due to structural changes and reorientation<br />

of particles at a given shear rate, the shear stress becomes time-dependent as the particles<br />

realign themselves to the flow. In other words, the shear stress takes time to readjust<br />

to the prevailing shear rate. Some of these changes may be reversible when the rate of reformation<br />

is the same as the rate of decay. However, in the case of flows in which the deformation<br />

is extremely slow, the structural changes or particle reorientation may be irreversible<br />

(see Figure 3-13).<br />

3-5-1 Thixotropic Mixtures<br />

When the shear stress of a fluid decreases with the duration of shear strain, the fluid is<br />

called thixotropic. The change is then classified as reversible and structural decay is observed<br />

with time under constant shear rate. Certain thixotropic mixtures exhibit aspects of<br />

permanent deformation and are called false thixotropic.<br />

When the rate of structural reformation exceeds the rate of decay under a constant sustained<br />

shear rate, the behavior is classified as rheopexy (or negative thixotropy).<br />

One typical example of a thixotropic mixture is a water suspension of bentonitic<br />

clays. These difficult slurries are produced by mud drilling associated with the use of<br />

positive displacement diaphragm or hose pumps. The reader may find throughout literature<br />

considerable discussion about “hysterisis.” This function is used to measure the<br />

behavior of the mixture by gradually increasing the shear rate and then by decreasing it<br />

back in steps. These curves are interesting but are of limited help to the designer of a<br />

pumping system.<br />

Shear Stress ( )<br />

Thixotropic<br />

Rheopectic<br />

Rate of shear ( = du/dy)<br />

FIGURE 3-13 Rheology of time-dependent fluids.


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

3.31<br />

Moore (1959) proposed expressing the complex behavior of a thixotropic fluid that<br />

does not possess a yield stress value in terms of six parameters:<br />

� = (�0 + c�)(d�/dt)<br />

d�/d� = a – �(a + bd�/dt)<br />

where<br />

� = duration of the shear for a time-dependent fluid<br />

a, b, c, and �0 = materials constants<br />

� = a structural parameter that has two values (0 and 1) at the limits where<br />

the material is fully broken down or fully developed<br />

Fredrickson (1970) discussed the modeling of thixotropic mixtures of suspensions of<br />

solids in viscous liquids and proposed that rheological tests be conducted to measure four<br />

constants to understand the qualitative nature of the mixture.<br />

Ritter and Govier (1970) proposed representing the behavior of thixotropic fluids as<br />

follows:<br />

� The formation of structures, networks, or agglomerates is similar to a second-order<br />

chemical reaction.<br />

� The breakdown of the structure is similar to a series of consecutive first-order chemical<br />

reactions where formation is meant by behavior that is time-dependent, whereas the<br />

breakdown occurs when the viscosity of the fluid acts as a Newtonian mixture that is<br />

independent of both the shear rate and the duration of shear (Figure 3-14).<br />

Shear stress, +0.01, lb /ft<br />

-1 10<br />

8<br />

6<br />

4<br />

10<br />

4<br />

2<br />

2<br />

-2<br />

10<br />

Duration of<br />

shear, min<br />

0<br />

1<br />

10<br />

100<br />

100 1000<br />

-1<br />

Rate of Shear, d /dt + 10 in sec<br />

FIGURE 3-14 Rheology of Pembina crude oil at 44.5°F at constant duration of shear. (After<br />

Govier and Aziz, 1972.)


3.32 <strong>CHAPTER</strong> THREE<br />

Ritter and Govier (1970) therefore proposed to express the shear stress of the fluid in<br />

terms of structural stress � s and � �, a component of shearing stress due to the Newtonian<br />

component of the fluid:<br />

� = � s + � �<br />

(3-53)<br />

log� � = –KD� �log � – log K �s – �s� �s0 + �s� ��<br />

DR (3-54)<br />

where<br />

�s0, �s� = structural stresses at a given shear rate after zero and infinite duration of shear<br />

�s0 = �0 – �(d�/dt)<br />

�s� = �� – �(d�/dt)<br />

KD = a constant that is independent of shear rate but is related to the first-order structural<br />

decay process and is expressed in the minutes –1 .<br />

KDR = a dimensionless measure of the interaction between the network or structure decay<br />

and the reestablishment processes<br />

The coefficient KDR is evaluated as<br />

�<br />

KDR = (3-55)<br />

where �s1 is measured after a lapse of 1 minute. In Equations 3-54 and 3-55, KDR, KD, �s0, �s1, and �s� are determined from rheology tests.<br />

Kherfellah and Bekkour (1998) examined the thixotropy of suspensions of montmorillonite<br />

and bentonite clays. Montmorillonite clays are used as thickening agents for<br />

drilling fluids, paints, pesticides, cosmetics, pharmaceuticals, etc. Commercial bentonite<br />

suspensions exhibited thixotropic properties for concentrations higher than 6% by weight.<br />

Rheopectic or negative thixotropic mixtures are not common in mining and will not be<br />

examined in this chapter.<br />

2 s0 – �s1�s� ��<br />

�s1�s� – � 2 � 2 (� s0/�s�) – �s �s0 – �s� s�<br />

3-6 DRAG COEFFICIENT <strong>OF</strong> <strong>SOLIDS</strong><br />

SUSPENDED <strong>IN</strong> NON-NEWTONIAN FLOWS<br />

Some solids may be transported by highly viscous fluids in a non-Newtonian flow<br />

regime. One such example includes solids transported in the process of drilling a tunnel in<br />

a sandy soil rich with clay or bentonite. Other examples of solids suspended in non-Newtonian<br />

flows are energy slurries, which are mixtures of fine coal and crude oils. In such<br />

circumstances, the drag coefficient of the coarse components is of interest.<br />

Brown (1991) reviewed the literature for settling of solids in non-Newtonian flows,<br />

but cautioned that the studies have been limited to single particles. Considerably more research<br />

is needed in this field.<br />

3-7 MEASUREMENT <strong>OF</strong> RHEOLOGY<br />

In the proceeding sections of this chapter, the concepts of Newtonian and non-Newtonian<br />

fluids were explored. Measuring the viscosity of a slurry mixture is recommended for ho-


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

mogeneous flows, mixtures with a high concentration of particles, and for fibrous and<br />

flocculated slurries.<br />

Subsieve particles are defined as particles with an average diameter smaller than<br />

35–70 �m (depending on whose reference book you consult). Slurry flows with subsieve<br />

particles at a relatively high concentration by volume (C v � 30%) are strongly rheologydependent.<br />

Heterogeneous flows, flows without subsieve particles, or flows with subsieve<br />

particles at a very low concentrations, are not governed by the rheology of the slurry.<br />

Flocculation or the addition of flocculates in the process of mixing slurries tends to result<br />

in non-Newtonian rheology.<br />

Rheology in simple layman’s terms is the relationship between the shear stress and the<br />

shear rate of the slurry under laminar flow conditions. Although this relationship extends<br />

to transitional and turbulent flows, most tests are conducted in a laminar regime, often in<br />

tubes or between parallel plates.<br />

3-7-1 The Capillary-Tube Viscometer<br />

The purpose of the capillary tube viscometer is to measure the rheology of a laminar flow<br />

under controlled velocity conditions. Tubes are used in a range of diameters from 0.8–12<br />

mm (1/32–1/2 in). The length of the tube is accurately cut to account for entrance effects<br />

and end effects. Typically, the length may be as much as 1000 times the inner diameter.<br />

The capillary tube viscometer is used to plot the average rate versus the shear stress at<br />

the wall of the tube. This is called the pseudoshear diagram, as defined by the<br />

Mooney–Rabinovitch equation:<br />

(du/dr) w = �0.75 + 0.2 8<br />

d[ln(8V/Di)] � ���<br />

(3-56)<br />

Di<br />

d[ln(�P/4Li)]<br />

where<br />

(du/dr) w = rate of shear at the wall<br />

�P = pressure drop due to friction over a length Li of pipe of inner diameter Di V = average velocity of the flow<br />

d = derivative<br />

The data is then plotted on a logarithmic scale as per Figure 3-15.<br />

The use of capillary-like viscometers is complicated by the “effective slip” of non-<br />

Newtonian fluid-suspended material, which tends to move away from the wall, leaving an<br />

attached layer of liquid. The result is a reduction in the measurements of effective viscosity.<br />

Therefore, it is often recommended to conduct such tests in a number of tubes of different<br />

diameters.<br />

Measuring the pressure loss between two points well away from the entrance and end<br />

effects gives the shear stress at the wall as:<br />

�w = Ri�P/(2Li) (3-57)<br />

By considering that the velocity profile at a height y above the wall is a function of the<br />

shear stress we obtain<br />

–(du/dy) w = f (�)<br />

It may be possible to establish a relationship between the flow rate Q and the shear stress<br />

� as<br />

Q<br />

� �R 3<br />

1<br />

= � �w �3 � w 0<br />

3.33<br />

� 2 f (�)d� (3-58)


3.34 <strong>CHAPTER</strong> THREE<br />

8V<br />

D<br />

shear rate<br />

For a Newtonian flow:<br />

or � = � w/(8V/D i).<br />

For a Bingham flow:<br />

for � > � 0, where � 0 is the yield stress.<br />

The velocity profile is expressed as<br />

2V �w = � = � (3-59)<br />

Di 4�<br />

� = �(du/dr) w + � 0<br />

= = � �2 2V � �w (� – �0)d� (3-60)<br />

By integration of this equation and by multiplying by 4, the shear rate is derived as<br />

8V<br />

� DI<br />

100<br />

10<br />

1.0<br />

Q<br />

� �R 3<br />

0<br />

0<br />

� Di<br />

� w<br />

� �<br />

water<br />

Q<br />

� �R 3<br />

increasing tube diameter<br />

Shear Stress<br />

� � w 3<br />

4<br />

�<br />

3<br />

� 0<br />

� �w<br />

= � 1 – � � + � �� (3-61)<br />

Equation 3-61 is called the Buckingham equation. This equation cannot be solved<br />

without long iterations. Many engineers prefer to simplify the Buckingham equation by<br />

ignoring the term (� 0/� w) 4 , as this term is of negligible magnitude compared with the other<br />

terms:<br />

� �0<br />

1.0 10<br />

1<br />

�<br />

3<br />

D<br />

2<br />

D 1<br />

D 3<br />

4 � 0<br />

�4 � w<br />

D 4<br />

D P<br />

4 L<br />

FIGURE 3-15 Pseudoshear diagram of a non-Newtonian mixture tested in a capillary tube<br />

rheometer.


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

�w � 8V�/Di + 4/3�0 The modified equation is plotted in Figure 3-16.<br />

(3-62)<br />

For a pseudoplastic slurry or power law fluid, the shear stress is expressed by Equation<br />

3-43. By analogy with the method developed for a Bingham flow in a tube, the following<br />

equation is expressed:<br />

= = � � 2 (�/K) 1/n Q 2V<br />

�3 �R<br />

1<br />

�3 �w<br />

�<br />

d� (3-63)<br />

� Di<br />

or<br />

= � �w (3-64)<br />

0<br />

which once integrated is expressed as<br />

= � � (3-65)<br />

The effective viscosity is expressed as<br />

�e = �w/(8V/Di) = K(8V/Di) (n–1) [4n/(3n + 1)] n �<br />

1/n<br />

2V n � w<br />

� � �1/n Di 3n + 1 K<br />

(3-66)<br />

Unfortunately, Equation 3-66 is of no value when n < 1.0, which is the case for many<br />

power law slurries. It would mean that as the shear rate increases, the effective viscosity<br />

decreases to zero. This is contradictory to nature. For power law exponents smaller than<br />

1.0, alternative equipment should be used to measure rheology.<br />

It is tricky to avoid errors with the use of capillary effect viscometers. A particular<br />

source of errors is the end effect. At the entrance exit of the tube, contraction and expansion<br />

of the flow cause additional pressure losses.<br />

(3+1/n)<br />

Q<br />

� ��<br />

3<br />

1/n<br />

�R (3 + 1/n)K<br />

w<br />

Shear Stress<br />

Velocity profile<br />

w<br />

2 r 0<br />

� �0<br />

shear rate<br />

FIGURE 3-16 Pseudoshear diagram for a Bingham plastic.<br />

dV<br />

�<br />

dy<br />

dU<br />

dy<br />

3.35


3.36 <strong>CHAPTER</strong> THREE<br />

3-7-2 The Coaxial Cylinder Rotary Viscometer<br />

A more practical instrument to use when measuring rheology is the coaxial cylinder viscometer.<br />

In basic terms, it is a device used to measure the resistance or torque when rotating<br />

a cylinder in a viscous fluid (Figure 13-17). The moment of inertia in the cylinder is<br />

established by the manufacturer.<br />

The torque is due to the force the fluid exerts tangentially to the outside surface of the<br />

cylinder:<br />

T = 2�R0h�wR0 (3.67)<br />

where<br />

T = (surface area) (shear stress) (radius)<br />

R0 = outside radius of the rotating cylinder<br />

h = height of the cylinder<br />

�w = shear stress at the wall<br />

The shear stress at any radius r in the fluid can be expressed as<br />

T du<br />

�w = � = ��<br />

2 2�r h dy<br />

If the liquid is rotating at an angular velocity �, then<br />

(du/dy) w = –rd�/dr<br />

r<br />

R c<br />

scale to measure torque<br />

rotation of bob at<br />

speed<br />

R<br />

FIGURE 3-17 The rotating concentric viscometer.<br />

0<br />

slurry<br />

(3.68)


and<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

� = –�rd�/dr<br />

d�<br />

� dr<br />

=<br />

� �<br />

d� = � 0<br />

Rc R0 –T<br />

� 2�h�r<br />

–T<br />

� dr 3 2�h�r<br />

3.37<br />

or<br />

� = � – � (3.69)<br />

where Rc is the radius of the outside cylinder.<br />

2 This is known as the Margulus equation. It is obvious that R 0 can be related to the moment<br />

of inertia Ik of the rotating bob cup.<br />

Since for a Bingham slurry, the rate of shear is expressed as du/dr = (� – �0); the Margulus<br />

equation can be demonstrated as<br />

� = � – � – ln� � (3.70)<br />

This equation is known as the Reiner–Rivlin equation.<br />

For a Pseudoplastic:<br />

2 1/n � = n[T/(2�R 0hK)] [1 – (R0/Rc) 2/n T 1 1<br />

� � � 2 2<br />

4�h� R 0 R c<br />

T 1 1 �0 Rc � � � � � 2 2<br />

4�h� R 0 R c � R0<br />

] (3.71)<br />

At the wall:<br />

2 �w = T/(2�R b h) (3.72)<br />

A plot of log �w versus log � can be constructed. The slope gives the flow index n and, by<br />

substituting Equation 3-45, the value of K can be calculated.<br />

Heywood (1991) discussed errors with the use of rotating viscometers. Particular<br />

sources of errors are the end effects from both cylinders and the possible deformation of<br />

the laminar layer under the effect of high rotational speed. Heywood recommended the<br />

use of cylinders with a long length to diameter ratio. Wall slip effects can be detected by<br />

using cylinders of different radius but same length. The vendors of rheometers publish<br />

equations to correct for wall slip and end effects.<br />

One important problem about the use of rheometers is that they do not distinguish between<br />

Bingham and Carson slurries. This can lead to grave mistakes in the design of a<br />

pipeline. Certain slurries have a course of fractions that could also precipitate during a<br />

rheometer test. Unfortunately, this would give false readings. When there is doubt, the<br />

safest approach is to conduct a proper pump test in a loop.<br />

Whorlow (1992) published a book on rheological techniques that includes dynamic<br />

tests and wave propagation tests. In the appendix, he listed a number of rheological investigation<br />

equipment manufacturers. Some of the techniques apply more to polymers and<br />

are not relevant to our discussion. Dynamic vibration tests have been extended to fresh<br />

concrete (Teixera et al., 1998). Concord and Tassin (1998) described a method to use<br />

rheo-optics for the study of thixotropy in synthetic clay suspensions. A rheometer optical<br />

analyzer was used on laponite, a synthetic hectorite clay. Laponite was mixed with water<br />

and tests were conducted at various intervals for up to 100 days. Rheo-optics seems to be


3.38 <strong>CHAPTER</strong> THREE<br />

FIGURE 3-18 Stresstech rheometer, courtesy of ATS Rehosystems. The rheometer was developed<br />

for the pharmaceutical and cosmetics industries, where materials consistency may<br />

vary from fluid to solid.<br />

a new technique based on the ability of solids to reorient themselves by applying to them<br />

a negative electrical charge.<br />

3-8 CONCLUSION<br />

In this chapter, it was demonstrated that mixtures of solids and liquids are complex systems.<br />

The size of the particles, the diameter of the pipe, the interaction with other particles,<br />

the viscosity of the carrier, and the temperature of the flow all interact to yield Newtonian<br />

or non-Newtonian flows.<br />

In the next three chapters, the principles discussed in the present chapter will be applied<br />

to calculate the velocity of deposition, the critical velocity, the stratification ratio,<br />

and the friction loss in closed and open conduits for heterogeneous and homogeneous<br />

mixtures.<br />

3-9 NOMENCLATURE<br />

a The longest axis of a particle in Albertson’s model<br />

A Parameter used to express viscosity of non-Newtonian flows


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

3.39<br />

A0 Coefficient<br />

A1 Coefficient<br />

b Axis of a particle in Albertson’s model<br />

B Parameter used to express viscosity of non-Newtonian flows<br />

c The shortest axis of a particle in Albertson’s model<br />

C Parameter used to express viscosity of non-Newtonian flows<br />

CD Drag coefficient of an object moving in a fluid<br />

CDo Profile drag coefficient of an object moving in a fluid<br />

CL Lift coefficient of an object moving in a fluid<br />

CN CS Cv Cv� Cw da Coefficient based on equivalent Ren Coefficient based on equivalent Ren Concentration by volume of the solid particles in percent<br />

Maximum packing concentration of solids<br />

Concentration by weight of the solid particles in percent<br />

Diameter of a sphere with a surface area equal to the surface area of the irregularly<br />

shaped particle<br />

dapp Apparent particle diameter<br />

df Apparent flocculant diameter<br />

dg Sphere diameter<br />

dn Diameter of a sphere with a volume equal to the volume of the irregularly<br />

shaped particle in Albertson’s model<br />

d� Particle diameter<br />

D Drag force<br />

Di Tube or pipe inner diameter<br />

E Factor between Albertson and Clift shape factors<br />

f( ) Function of<br />

FBF Buoyancy force<br />

Fw Wall effect correction factor for free-fall speed of a particle<br />

g Acceleration due to gravity (9.78–9.81 m/s2 )<br />

gc Conversion factor, 32.2ft/s2 if U.S. units between lbms and slugs<br />

h Height of the cylinder<br />

Ik Moment of inertia<br />

K Consistency index or power law coefficient for a pseudoplastic<br />

KD A constant that is independent of shear rate but is related to the first-order<br />

structural decay process and is express in minutes –1<br />

KDR A dimensionless measure of the interaction between the network or structure<br />

decay and the reestablishment processes<br />

Kt Coefficient for terminal velocity<br />

Kz Kozney constant<br />

K1, K2, K3 Coefficients<br />

ln natural logarithm<br />

L Lift force<br />

Lc Characteristic length<br />

LI Length of pipe or tube<br />

n Flow behavior index, or exponent for a pseudoplastic (


3.40 <strong>CHAPTER</strong> THREE<br />

Ren Reynolds Number of a particle based on dn Rep Reynolds Number of a sphere particle based on its diameter<br />

Ri Inner radius of a pipe or tube<br />

R0 Radius of the bob in the coaxial cylinder rotary viscometer<br />

sp The surface area per unit volume of a sphere of equivalent dimensions or<br />

6/dg, also called specific surface of a particle<br />

Sf Front area of a particle orthogonal to the direction of flow<br />

Sw Surface area of a wing along the direction of flight<br />

T Applied torque for the cylinder rotary viscometer<br />

Ta Absolute temperature<br />

V Average velocity of the flow<br />

V0 Terminal velocity at very low volume concentration of solids<br />

Vc Terminal velocity at given volume concentration of solids<br />

Vt The terminal (or free settling) speed<br />

W Weight<br />

� the ratio of immobilized dispersing fluid to the volume of solids related approximately<br />

to the particle and floc apparent diameter<br />

� A coefficient used to express to the shear stability of a pseudoplastic mixture<br />

� Concentration by volume in decimal points<br />

� Shear strain<br />

d�/dt Wall shear rate or rate of shear strain with respect to time<br />

� Coefficient of rigidity of a non-Newtonian fluid, also called Bingham viscosity<br />

� A structural parameter for thixotropic fluids, which do not possess a yield<br />

stress value<br />

� Carrier liquid absolute viscosity<br />

�a Apparent viscosity of a pseudoplastic fluid<br />

�e Effective viscosity<br />

�0 Apparent viscosity of a pseudoplastic fluid at zero shear rate<br />

�� Bingham plastic limiting viscosity, or apparent viscosity of a pseudoplastic<br />

fluid at very high shear rate<br />

� Pythagoras number (ratio of circumference of a circle to its diameter)<br />

� Duration of the shear for a time-dependent fluid<br />

� Density<br />

� Shear stress at a height y or at a radius r<br />

�0 Yield stress for a Bingham plastic or yield pseudoplastic<br />

�s Structural stress of a thixotropic fluid<br />

�w Wall shear stress<br />

� Kinematic viscosity<br />

� Angular velocity of particle<br />

� Angular velocity of complete system<br />

� The logarithmic standard deviation<br />

�A Albertson shape factor<br />

�c Clift shape factor<br />

Thomas shape factor<br />

� T<br />

Subscripts<br />

g Equivalent sphere<br />

L Liquid<br />

m Mixture<br />

p Particle<br />

s Solids


3–10 REFERENCES<br />

<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

3.41<br />

Albertson, M. L. 1953. Effects of shape on the fall velocity of gravel particles. Paper read at the 5th<br />

Iowa Hydraulic Conference, Iowa University, Iowa City, Iowa.<br />

Allen, H. S. 1900. The motion of a sphere in a viscous fluid. Phil. Mag., 50, 323–338, 519–534.<br />

Boger, D. V., and Q. D. Nguyen. 1987. The Flow Properties of Weipa #3 and #4 Plant Tailings. Internal<br />

study conducted by Comalco Aluminium Ltd, Weipa, Australia, quoted in Darby, R., R.<br />

Mun, and D. V. Boger. 1992. Predict Friction Loss in Slurry Pipes. Chem. Engineering, 99, 9<br />

(September), 117–211.<br />

Brown, G. G. 1950. Unit Operations. New York: Wiley.<br />

Brown, N. P. 1991. The settling behavior of particles in fluids. In Slurry Handling, Edited by N. P.<br />

Brown and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Caldwell, D. H., and H. E. Babitt. 1941. Flow of muds, sludge and suspensions in circular pipe. Am.<br />

Inst. Chem. Engrs. Trans., 37, 2 (April 25), 237–266.<br />

Caldwell, D. H., and H. E. Babitt. 1942. Pipeline flow of solids in suspension. Symp. Am. Soc. of Civ.<br />

Eng. Proc., 68, 3 (March), 480–482.<br />

Cheng, D. C. H., and W. Whitaker. 1972. Applications of the Warren Spring Laboratory pipeline design<br />

method to settling suspension. Paper read at the 2nd Annual Hydrotransport Conference,<br />

Bedford, England.<br />

Chilton, R. A., and R. Stainsby. 1998. Pressure loss equations for laminar and turbulent non-Newtonian<br />

pipe flow. Journal of Hydraulic Engineering, 124, 5 (May), 522–529.<br />

Clift, R., J. R. Grace, and M. E. Weber. 1978. Bubbles, Drops and Particles. New York: Academic<br />

Press.<br />

Concord, S., and J. F. Tassin. 1998. Rheoptical study of thixotropy in synthetic clay suspensions. In<br />

Proceedings of the Fifth European Rheology Conference. Ljubljana, Slovenia: University of<br />

Ljubljana.<br />

Cross, M. M. 1965. Rheology of non-Newtonian fluids—New flow equation for pseudoplastic systems.<br />

Journal of Colloid Science, 20, 417.<br />

Dahlgreen, S. E. 1958. Eyring model of flow applied to thixotropic equilibrium. Journal of Colloid<br />

Science, 13 (April), 151–158.<br />

Dedegil, M. Y. 1987. Drag coefficient and settling velocity of particles in non-Newtonian suspensions.<br />

Journal of Fluids Engineering, 109 (September), 319–323.<br />

Dick, R. I., and B. B. Ewing. 1967. Rheology of activated sludge. Journal of Water Pollution Control<br />

Federation, 39, 543.<br />

Dahlgreen, S. E. 1958. Eyring model of flow applied to thixotropic equilibrium. Journal of Colloid<br />

Science, 13 (April), 151–158.<br />

Fredrickson, A. G. 1970. A model for the thixotropy of suspensions. American Inst. of Chem. Eng.<br />

Journal, 16, 436.<br />

Gay E. D., P. A. Nelson, and W. P. Armstrong. 1969. Flow properties of suspensions with high<br />

solids concentration. American Inst. of Chem. Eng. Journal, 15, 6, 815–822.<br />

Goodrich and Porter. 1967.<br />

Govier, G. W., C. A. Shook, and E. O. Lilge. 1957. Rheological Properties of water suspensions of<br />

finely subdivided magnetite, galena, and ferrosilicon. Trans. Can. Inst. Mining Met., 60, 157.<br />

Govier, G. W., and K. Aziz. 1972. The Flow of Complex Mixtures in Pipes. New York: Van Nostrand<br />

Reinhold.<br />

Hedstrom, B. O. A. 1952. Flow of plastic materials in pipes. Ind. Eng. Chem., 33, 651–656.<br />

Herbrich, J. 1968. Deep ocean mineral recovery. Paper read at the World Dredging Conference II,<br />

Rotterdam, the Netherlands.<br />

Herbrich, J. 1991. Handbook of Dredging Engineering. New York: McGraw-Hill.<br />

Heywood, N. I. 1991. Rheological characterisation of non-settling slurries. In Slurry Handling, Edited<br />

by N. P. Brown and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Heywood, N. I. 1996. The performance of commercially available Coriolis mass flowmeters applied<br />

to industrial slurries. Paper read at the 13th International Hydrotransport Symposium on<br />

Slurry Handling and Pipeline Transport. Johannesburg, South Africa. Cranfield, UK: BHRA<br />

Group.<br />

Inter-Agency Committee on Water Resources. 1958. Report 12. Internal report by the Subcommittee<br />

on Sedimentation, Minneapolis, Minnesota.


3.42 <strong>CHAPTER</strong> THREE<br />

Kearsey, H. A., and L. E. Gill. 1963. Study of sedimentation of flocculated thorium slurries using<br />

gamma ray technique. Trans. Inst. Chem. Engrs., 41, 296.<br />

Kherfellah, N., and K. Bekkour. 1998. Rheological characteristics of clay suspensions. In Proceedings<br />

of the Fifth European Rheology Conference. Ljubljana, Slovenia: University of Ljubljana.<br />

Krusteva, E. 1998. Viscosmetric and pipe flow of inorganic waste slurries. In Proceedings of the<br />

Fifth European Rheology Conference. Ljubljana, Slovenia: University of Ljubljana.<br />

Lapassin, R., S. Pricl, and M. Stoffa. 1998. Viscosity of aqueous suspensions of binary and ternary<br />

alumina mixtures. In Proceedings of the Fifth European Rheology Conference. Ljubljana,<br />

Slovenia: University of Ljubljana.<br />

Metzner, A. B., and M. Whitlock. 1958. Flow behavior of concentrated (dilatant) suspensions. Trans.<br />

Soc. Rheology, 2, 239–254.<br />

Moore, F. 1959. Rheology of Ceramic Slips and Bodies. British Ceramic Society Transactions, 58,<br />

470.<br />

Mun, R. 1988. The Pipeline Transportation of Suspensions with a Yield Stress. Master’s Thesis, University<br />

of Melbourne, Australia.<br />

Parzonka, W. 1964. Determination of the maximum concentration of homogeneous mixtures (in<br />

French). Journal of the French Academy of Science, 259, 2073.<br />

Pilpel, N. 1965. Flow properties of non-cohesive powders. Chemical Process Eng. 46, 4, 167–179.<br />

Prokunin, A. N. 1998. Particle-wall interaction in liquids with different rheology. In Proceedings of<br />

the Fifth European Rheology Conference. Ljubljana, Slovenia: University of Ljubljana.<br />

Richards, R. H. 1908. Velocity of Galena and Quartz Falling in Water. Trans AIME, 38, 230–234.<br />

Ritter, R. A., and G. W. Govier. 1970. The development and evaluation of a theory of thixotropic behavior.<br />

Can. Journal Chem. Eng., 48, 505.<br />

Rubey, W. W. 1933. Settling velocities of gravel, sand and silt particles. Amer. Journal of Science,<br />

25, 148, 325–338.<br />

Skelland, A. H. P. 1967. Non-Newtonian Flow and Heat Transfer. New York: Wiley.<br />

Teixeira, M. A. O. M., R. J. M. Craik, and P. F. G. Banfill. 1998. The effect of wave forms on the vibrational<br />

processing of fresh concrete. In Proceedings of the Fifth European Rheology Conference.<br />

Ljubljana, Slovenia: University of Ljubljana.<br />

Thomas, D. G. 1961. Transport characteristics of suspensions: Part II. Minimum transport velocity<br />

for flocculated suspensions in horizontal pipes. AIChE Journal, 7 (September), 423–430.<br />

Thomas, D. G. 1963. Transport characteristics of suspensions. Ch. E. Journal, 9, 310.<br />

Thomas, A. D. 1981. Slurry pipeline rheology. Paper presented at the National Conference on Rheology.<br />

Second Annual Conference of the British Society of Rheology, Australian Branch, University<br />

of Sydney, Australia.<br />

Turton, R., and O. Levenspiel. 1986. A short note on drag correlation for spheres. Powder Technology<br />

Journal, 47, 83.<br />

Valentik, L., and R. L. Whitemore. 1965. Terminal velocity of spheres in Bingham plastics. British<br />

Journal of Applied Phys., 16, 1197.<br />

Vlasak, P., Z. Chara, and P. Stern. 1998. The effect of additives on flow behaviour of kaolin–water<br />

mixtures. In Proceedings of the Fifth European Rheology Conference. Ljubljana, Slovenia:<br />

University of Ljubljana.<br />

Wasp, E. J., J. P. Kenny, and R. L. Gandhi. 1977. Solid-Liquid Flow—Slurry Pipeline Transportation.<br />

Trans-Tech Publications.<br />

Wells, P. J. 1991. Pumping non-Newtonian slurries. Technical Bulletin 14. Sydney, Australia: Warman<br />

International.<br />

Whorlow, R. W. 1992. Rheological Techniques, 2d. ed. New York: Ellis Horwood.<br />

Wilson, K. C., G. R. Addie, and R. Clift. 1992. Slurry Transport Using Centrifugal Pumps. New<br />

York: Elsevier Applied Sciences.<br />

Worster, R. C., and D. E. Denny. 1955. Hydraulic transport of solid materials in pipes. Proceedings<br />

of the Institute of Mechanical Engineers (UK), 38, 230–234.<br />

Further Reading:<br />

Caldwell, D. H., and H. E. Babitt. 1942. Pipeline flow of solids in suspension. Symp. Am. Soc. of Civ.<br />

Eng. Proc., 68, 3 (March), 480–482.<br />

Goodrich, J. E., and R. S. Porter. 1967. Rheological interpretation of torque—Rheometer data. Polymer<br />

Eng & Science, 7 (January), 45–51.


<strong>MECHANICS</strong> <strong>OF</strong> <strong>SUSPENSION</strong> <strong>OF</strong> <strong>SOLIDS</strong> <strong>IN</strong> <strong>LIQUIDS</strong><br />

3.43<br />

Lazerus, J. H., and P. T. Slatter. 1988. A method for the rheological characterization of tube viscometer<br />

data. Journal of Pipelines, 7, 165–176.<br />

Thomas, D. G. 1960. Heat and momentum transport characteristics of non-Newtonian aqueous thorium<br />

oxide. AIChE Journal, 7, 431.<br />

Wilson, K. C. 1991. Pipeline design for settling slurries. In Slurry Handling. Edited by N. P. Brown,<br />

and N. I. Heywood. New York: Elsevier Applied Sciences.<br />

Wilson, K. C. 1991. Slurry transport in flumes. In Slurry Handling. Edited by N. P. Brown, and N. I.<br />

Heywood. New York: Elsevier Applied Sciences.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!