Вы находитесь на странице: 1из 14

ELSEVIER

Int. J. Miner. Process. 53 (1998) 15-28

minERAL PROIESSInG

RIIERIIIITIOIR_Of

Spirals observed
P.C. Kapur
a

a,

*,

T.P. Meloy

Department of Materials and Metallurgical


b

Engineering, Indian Institute of Technology, Kanpur 208016, India Particle Analysis Center, 338 Comer, West Virginia University, Morgantown, WV 26506, USA

Accepted 15 Apri11997

Abstract We first examine the three main components required to construct a realistic and tractable model of the working of a spiral, namely (1) geometry of the spiral and its trough, (2) principal forces acting on a particle which are gravity, centrifugal, hydrodynamic drag and lift and friction forces, and (3) flow of fluid down the curvilinear tightly turning path of the spiral. Next we combine these elements seamlessly by assuming that the particles eventually attain dynamic equilibrium in the forward longitudinal direction and static equilibrium in the transverse direction. The resulting force function provides a spectrum of the particles' radial location on the trough according to their size and density. Simulation results are highly encouraging. The model faithfully reflects the present empirical spiral design philosophy for treating different feeds such as coal or heavy density minerals. It seemingly has potential for computer-aided spiral design and in control and optimization of spiral circuits. 1998 Elsevier Science B. V.
Keywords:

spirals; gravity separators; coal cleaning; modeling and design

1. Introduction After suffering a benign neglect for nearly half a century, spirals have begun to attract somewhat greater attention of the researchers in recent years. The reason presumably lies in the world-wide concern with the environmental damage which is being wrought by the combustion of fossil fuels. Spirals are perceived as rugged, compact, efficient and cost-effective gravity separators for cleaning coal in the size
Corresponding author. Present address: Tata Research Development and Design Centre 54 B, Hadapsar Industrial Estate, Pune 411 013, India. E-mail: pckapur@deI2.vsnl.net.in 0301-7515/98/$19.00 1998 Elsevier Science B.V. All rights reserved. PII S0301-7516(97)00053-7

16

P.e. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 (1998) 15-28

range of 0.1 nun to 2-3 mm (Burt, 1984; Palowitch et al., 1991; Mankosa et al., 1995; Tavares and Sampaio, 1995). The basic contours of the spiral design have hardly changed since their introduction in the 1940s. Only limited and incremental modifications have been incorporated from time to time by trial and error procedures and experience gained from plant operation. In other words, our understanding of the working of a modem spiral is mostly empirical. We present a preliminary theoretical analysis of spirals for developing a relatively tractable mathematical model that hopefully mimics the performance of this equipment with a reasonable degree of accuracy and realism. As pointed out by many authors, such a model could provide a valuable simulation tool for the design and scale up of this device and for the control and optimization of spiral circuits (Sivamohan and Forssberg, 1985; Holland-Batt, 1989; Holland-Batt and Holtham, 1991; Loveday and Cilliers, 1994; Li et al., 1995). It is reported that extensive analysis and modeling of spirals have been carried out in the former Soviet Union (Burt, 1984), but details of this body of work is not available to us. A fairly elaborate and sophisticated spiral model was proposed by Holland-Batt (1989). The model, however, makes heavy computational demands in calculating the local fluid flows and particle velocities. Moreover, it assumes a free motion regime for particles which is difficult to reconcile with the experimental measurements, especially in the inner region of the spiral trough where solids concentrations in the slurry can rise as high as 75% or more. Recently, Li et al. (1995) presented the results of their force balance spiral model without providing any information on the structure of the model. There are three broad components of a spiral which must be integrated into any plausible spiral model. These are: (1) spiral geometry; (2) forces acting on particles moving down the spiral and their magnitudes; and (3) flow characteristics of the fluid (strictly speaking, the slurry), which in tum affects, with the exception of gravity force, all other forces either directly or indirectly. We next proceed to discuss these components by tum.

2. Spiral geometry It is evident that the performance of a spiral would depend strongly on its design parameters, including diameter, height, number of turns, pitch and slope as well as the shape of the channel or trough and its dimensions. Geometrically, the deck of the channel may be visualized as comprising an infinitely large number of axially adjacent non-intersecting helical curves. The parametric equations of an helix in x-y-z Cartesian coordinates are given by (von Seggem, 1990):
x
y
=

r sin [t]

(1)

= r cos [ t] ; 0 s t 27T '

s N7T

(2) (3)

u z= -to O::s;z::s;H

P.e. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 (1998) 15-28 Centre Column

17

Height, H

Outer Radius, 'o

,'.--+' ,
~,-l

Max. DePth_.-_~_~ -_-_- -:, ......

Trough Slope

Inner Radius, ri

Fig. 1. Schematic drawing of a spiral and its trough.

where N is twice the number of turns, H is spiral height and r is radial distance from the centerline. From these relationships the longitudinal tangential slope S at any point on the deck may be derived as: S=tan[a] =u 27Tr

(4)

where u is pitch of the spiral and a is slope angle shown in Fig. 1. Most trough shapes are of the form of a modified quarter-circular arc (Burt, 1984). Accordingly, these may be represented, exactly or approximately, by an elliptical curve, also shown in Fig. 1. For this, shape the local slope of the channel deck in radial or transverse direction is given by:

(5)
where 1I is local slope angle in the transverse direction, c y is the maximum depth of the trough and rj and ro are, respectively, the inner and outer radii of the trough from the centerline. Our choice of the elliptical contour for the trough is dictated primarily by mathematical convenience of realizing analytical or simpler expressions. We are of

18

P.e. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 (1998) 15-28

course free to choose any arbitrary contour without introducing any new principle in the development of our model.

3. Forces acting on particles

Because of the presence of mixed laminar, transitional and turbulent flow regimes, particles on the spiral are subjected to a medley of randomly fluctuating, transient and steady force fields. It is not easy to identify and quantify most of these forces precisely. In general only rough estimates are possible of the five principal forces involved, namely, gravity, centrifugal, hydrodynamic drag, lift and friction forces.
3.1. Gravity force

Force of gravity or weight of a spherical particle of diameter d and density submerged in a liquid of density p is given by:

(T

(6)
where
g

is the gravitational constant.

3.2. Centrifugal force

In view of the fast flowing and tightly turning flow obtaining in a spiral, centrifugal force plays a significant role in its operation. A submerged particle moving with a velocity v in a circular path of radius r is subjected to a centrifugal force of:
7T

Fc=------

d3 v 2 (

(T -

p)

(7)

In general, velocity of the particle will be different from that of the surrounding fluid, as well as highly variable depending upon its size, density, position in the radial direction and depth under water. A theoretical analysis of the particle velocity in a rotating two-phase flow is not a trivial task. Consequently, we are obliged to adopt a force balance strategy for surmounting this problem, as elaborated in a later section.
3.3. Drag force

The drag force exerted on a body in a flowing stream of fluid due to skin friction and eddies induced pressure difference between its upstream and downstream sides is given by (Allen, 1982a):
Fd =

'4 p ghd?

7T

sin] a]

(8)

where

is depth of flow.

P.C. Kapur, T.P. Meloy / Int. 1. Miner. Process. 53 (1998) 15-28

19

3.4. Lift force Large-size particles in the vicinity of the channel deck are subjected to hydrodynamic lift (Bagnold) forces that arise by the influences of nearby boundary, fluid shear and particle spin or the Magnus effect (Bagnold, 1954, 1966; Allen, 1982a). In spirals lift forces assist in dilating and loosening the bed which facilitates segregation and sorting of particles according to their size and density. As a first approximation lift forces can be directly related to drag force as:

(9)
At present no consensus exists on the constant of proportionality k1; experimentally determined values ranging from 1/7 to unity have been reported (Simon and Senturk, 1976). Since the spiral simulation results are not overly sensitive to its value, we arbitrarily set kl = 0.33. 3.5. Friction forces Resistance to the motion of the particle on spiral trough is proportional to the sum total of the normal components of all forces FN acting on the particle. The constant of proportionality is the coefficient of dynamic friction under water tan[j] which, according to Bagnold, is virtually the same as the coefficient of static friction under water (Allen, 1982a). Hence friction force is related to normal force by:

(10)
where based on the data provided by Allen and others (Chang. 1988), we take tan[ cf>] = 0.5. All gravity separation processes in essence entail countering the force of gravity on particles by imposing some other forcers). Extensive computations of these forces (Kapur and Meloy. 1996) reveal the following noteworthy points. One, with the exception of gravity force, other forces are strongly dependent on the spiral geometry. Two, the magnitude of gravity force is comparable to that of drag and centrifugal forces in size ranges that are normally processed in spirals. And three, it is not so much the magnitudle of individual forces as the rate of change of forces with particle size, density and radial position which determines the efficiency of separation on spirals.

4. Fluid flow on spirals Of all the different kinds of gravity separators available, spirals exhibit the most complex flow regime. The action of centrifugal force on water as it flows down in a spiral channel has two important consequences (Allen, 1982b; Chang, 1988). First, the water level at the outer concave wall of the trough exceeds that at the inner convex surface by an amount called superelevation. Second. a transverse secondary circulation is generated in the form of a vertically flattened helical spiral that moves forward in a

20

P.e. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 (1998) 15-28

corkscrew fashion. The angles that the inward and outward bound flows makes with the mean axial flow vary with depth and radial distance. Holtham (1990, 1992) and Holland-Batt and Holtham (1991) have measured these angles in water only spirals with rather large error bounds. It is however not clear if angles obtained in pure water are meaningful for heavily loaded slurries. We find it convenient to utilize a widely quoted expression in the hydrology literature by Rozovskii for the mean deviation angle 8 (Allen, I982b): tan[ Analysis equations a matter precisely described

8] = 11-

h
r

(11 )

of fluid velocity and local flow profiles in a spiral by the Navier-Stokes is an arduous task indeed. Consequently, we adopt an 'averaging' approach as of practical expediency and with no pretentions for describing the flow either or in detail. Flow of fluids and sediments in open channels are commonly by a plethora of power laws whose general form is: ( 12)

where V is mean flow velocity, R is hydraulic radius and K is a composite resistance coefficient. The exponents a and b depend on the nature of the flow such as laminar, Manning laminar, Lacey rough channel, Blasius turbulent, Bagnold suspension, transitional or mixed type or a combination thereof (Dingman, 1984; Chen, 1991). We choose the transitional or mixed flow equation because it is widely employed to describe the sheet or overland flows of water during rainstorm where energy dissipation occurs, as it does in spirals, by viscosity and turbulence frictions combined in relatively shallow depths. Moreover, the mean flow velocities computed with this equation are comparable with the measured data, even though the computed flow depths tum out to be greater than the reported values. The power law for transitional or mixed flow is described as: 26.4 V= _-RS1/2 dlp /6

(13)

where dp is a percentile size in the solid feed. Another expression for the mean flow velocity is: Q V=A (14)

where Q is volumetric feed rate and A is cross-sectional area of flow. Next, we set Sin Eq. (13) equal to Sm' the channel tangential slope at the midpoint and substitute appropriate expressions for R and A in terms of the trough geometry and mean flow depth hm and finally eliminate V between Eqs. (13) and (14) to arrive at:

Q-

_ 3.377' 1.5 2 dl/6


p

fE
ex

_'__~'1T-:-;/2:;--;:::======::::-''----~
V(xI2+yI2)dt+hm

[ex e y - ( ex - hm) ( e y - hm) ] 2

1 o

(15)

P.C. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 (J998) 15-28

21

where x' and y' are derivatives of x and y with respect to t and ex = To - Tj is the radial width of the trough. Eq. (15) can be solved for mean flow depth which is required in Eqs. (8) and (I1) for implementing the force balance model.

5. Force eqmlibrium model


Fluid driven particles can travel by one or more of the following modes: sliding, rolling, saltation (i.e, bounding), known collectively as the bedload transport, and in suspension. Considering the few rom thick rotatory pulp flow obtained in a spiral, it is reasonable to expect that, barring low density fines, movement of particles is by a combination of mechanisms in which the bedload transport should play a dominant role. In any case, it is not possible to track the trajectory of individual particles. We can at best monitor the overall 'drift' by making some plausible assumptions. If and when a steady state is attained, particles are in dynamic equilibrium in the longitudinal direction and in static equilibrium in the transverse direction. The latter state is of primary interest as it determines the degree and extent of sorting of particles according to their density

---------(------Direction

Spiral Slope Angle

Transverse

"' ...... "" 6


Direction

D*a

II

Fig. 2. Forces acting on a particle lying on the spiral trough. showing longitudinal and the local shear deviation angle.

and transverse slope angles

22

P.e. Kapur, T.P. Meloy

I Int.

J. Miner. Process. 53 (1998) 15-28

and size. As seen in Fig. 2, the longitudinal component of all forces acting on a particle in steady motion is:

FL = Fg sin[ ()] sin] a] - F;, cos[ ()]sin[ a]


where the normal component of forces is:

+ Fd cos[ S] - FN tan[

c/>]

=0

(16)

FN = Fgcos[ ()]

+ Fe sin[ ()]

- FJ

( 17)

The transverse component of forces acting on a stationary particle is:

FT = Fe cos[ ()]cos[

a] + e, sin[ S] - Fg

sin[ ()]cos[

a] = 0

( 18) force term yields a

Combining these expressions force function given below: FJtan[ c/>]

with the elimination

of centrifugal

__:_-------=-----=-----=-~----::----=-----::--::::-----.::--~ cos[ e ]sin[ a] + tan[ c/> ]sin[ ()]


(Fdsin[ ()] - Fgcos[ a ]sin[ ()])

+ FdCOS[ ()] - Fg(

tan [ c/> ]cos[ ()] - sin[ a ]sin[ (}])

+ sec[

a ]sec[ ()]

=0

(19)

Substitutions for gravity, drag and lift forces given in Eqs. (6), (8) and (9) yield: phS 4Ji+S2 _ dg tan[ c/> ]( a - p) [ 6 cos[ [ kJtan[ c/> ]

+ cos[

()]

+ sin[

() ]tan[ a ] + tan[ c/> ]sec[ a ]sin[ S ]tan[ ()]

cos[ ()]sin[ cos[ ()]

a] + tan[ c/> ]sin[ ()]


()] ]

e ]sin[

+ sin[ ()]tan[

=0
(20)

a] + tan[ c/> ]sin[ ()]

Substitutions for S and angles a, () and () given in Eqs. (4), (5) and (11) and replacement of h by hm from Eq. (15) results in a rather complicated and long

Fig. 3. Simulated 3-D plot of particle relative density as a function of size and equilibrium radial position on the trough for a spiral of specifications ri = 0.1 m, ro = 0.325 m, cy = 0.15 m and u = 0.35 with a slurry feed rate of 4 m3/h.

P.e. Kapur, T.P. Meloy / Int. 1. Miner. Process. 53 (J998) 15-28

23

expression which for our purposes may be written in the form of a force equilibrium function as follows: (21) Using this function we can compute, after inserting 'best' estimates of k, and cp available in the literature, the equilibrium radial position r on the trough of a particle of size d and density (T for any spiral of design specifications r ro' C y and u.
j,

6. Simulation results

The force equilibrium function in Eq. (21) can be readily solved for particle density as a function of particle size and radial position or for particle size as a function of

Cii

II:

1.5~--+-=rr:==1P;;;;::::::=f=~~~

1~------_L------~------~------~~~~
100
150 200 250
Eq. Position, r mm

300

E E
'0

i:i5

100

150

200
Eq. Position, r mm

250

300

Fig. 4. Parametric plots of simulation shown in Fig. 3.

24

P.e. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 (1998) 15-28

<-chromite Silica-> 0.8 ~--'~---,+----'''''=1-;----T-----j

D.6~--~1-~~~q--~~+----~

.g
U t:
til

~0.4~---~----~-~~+--~-~
0..0.2

100

150

200

250

300

Equilibrium Position, r [mm]


Fig. 5. Comparison of simulated results and experimental data for 0.85 XO.71, 0.71 XO.6 and 0.6XO.5 mm size fractions of chromite and silica separated on a Multotec '21 degree' heavy mineral spiral (Loveday and Cilliers, 1994).

particle density and radial position. Fig. 3 shows a 3-D plot of relative density for a spiral of rj = 0.1 m, ro = 0.325 m, cy = 0.15 m and u = 0.35 with a slurry feed rate of 4 m3/h. As observed in the actual spiral operation, here also the lighter particles in the simulation are segregated in the outer region while heavier particles are concentrated in the inner region of the trough. These results are perhaps more clearly illustrated in the parametric plots in Fig. 4 of particle size or particle density versus the equilibrium radial position. It is rather remarkable that the force equilibrium spiral model with its fairly complicated expressions and without any fine tuning of the parameters culled from the hydrology literature, nevertheless, yields realistic values for both particle size and particle density. Loveday and Cilliers (1994) separated chromite from silica on a Multotec '21 degree' heavy mineral spiral of specifications rj = 0.1 m, ro = 0.3 m, cy = 0.125 m and u = 0.45 with a slurry feed rate of 3.5 m3/h. The simulated results and experimental data for 0.85 X 0.71,0.71 X 0.6 and 0.6 X 0.5 mm size fractions are compared in Fig. 5. Li et al. (1995) measured the classification of color-coded plastic particles of different densities on a LD2 spiral of specifications rj = 0.075 m, ro = 0.33 m, cy = 0.135 m and u = 0.308 with a slurry feed rate of 3 m3/h. Their data for three size fractions is compared with the model results in Fig. 6. Considering the complexity of the spiral system, the simplifications that we are obliged to introduce in the model, and the unavoidable presence of not insignificant levels of errors in the measurements, we may conclude that the model is able to mimic the spiral behavior reasonably well. Over the years, some sort of empirical consensus has evolved on the design of spirals for a given separation task. The spiral specifications for coal cleaning, for example, according to Palowitch et al. (1991), are 0.6 m diameter and 0.254 m pitch with five and half turns. Holland-Batt and Holtham (1991) provided a table of different spirals in which coal spiral LD2 has 0.33 m trough radius and 0.308 m pitch while high ash coal spiral LD9 has 0.35 m trough radius and 0.273 m pitch. Earlier, Burt (1984) listed

P.e. Kapur, T.P. Meloy / Int. 1. Miner. Process. 53 (J998) 15-28


Feed Size 1xO.71 mm

25

1-

-. <,
~
r-...
2
1.6 Relative

1.4

1.8
Density

---

2.2

Feed Size 0.51xO.355 ~50 ';:'300 C

mm

r-....

.............

o a.
:J

~250 II) E200

J--_
I

CT W

@150 '5 100

1.4
1.6 Relative

2.2

1.8
Density

Feed Size 0.09xO.063

mm

1.4
1.6 Relative

1.8
Density

2.2

Fig. 6. Comparison of simulated results and experimental spiral et aI., 1995).

ru

data for color-coded

plastics separated on a LD2

Reichert spirals Mark 9 and 10 developed specifically for coal which also have a pitch of 0.273 m. In contrast, heavy minerals require a significantly larger pitch of about 0.4 to 0.45 m with somewhat smaller diameter (Holland-Batt and Holtham, 1991; Loveday and Cilliers, 1994). To be meaningful as a simulation tool, our spiral model must also reflect these design trends. Fig. 7 shows the simulated effect of spiral pitch on 1 mm size coal feed. The pitch varies from 0.25 m (bottom curve) to 0.5 m (top curve) in increments of 0.05 m. The negative slope of the curves, that is, density drop per unit radial distance, is a measure of the separation efficiency. In other words, the lower the

26

P.e. Kapur, T.P. Meloy / Int. 1. Miner. Process. 53 (1998) 15-28


ParticleSize 1 mm

sr-------,-------,-------~------~--,
4.S ~---__+----+___-4~---__+----+___---_t_-->--t_----+------1

~3.S~----~----t-------t-----T-~
Q)

~ 3~~~~__+----i-----_t_---_T-~

~ w2.S~~~~~~~~-t_---_t_---_T-~
c::

>

1~------~-------~----~------~----~
100 1S0 200 2S0 300
Equilibrium Position, r [mm]

Fig. 7. Effect of spiral pitch on the separation of I mm size particles in a spiral of specifications rj = 0.1 m, ro = 0.325 m and cy = 0.15 m. Pitch varies from 0.25 m for the bottom curve to 0.5 m for the top curve with an increment of 0.05 m.

slope, the greater is the efficiency. It will be seen that the optimum pitch for feed in the range of 1.2 to 2.8 relative density lies between 0.25 and 0.3 m, which is the same as the current industrial practice. Fig. 8 shows the effect of trough radius where ro is varied from 0.25 m (left most curve) to 0.5 m (right most curve) in increments of 0.05 m. Again, it would seem that spirals with larger radii should exhibit superior performance in coal cleaning.

ParticleSize

1 mm

S
4.S

~
Q)

~3.S
3

.~

:m2.5

c::

~ ~ ~

2 1.5 1

I'-.200

'\ j ",--........,
300
Position, r [mm]

,~
SOO

100

400

Equilibrium

Fig. 8. Effect of trough outer radius on the separation of 1 mm size particles in a spiral of specifications rj = 0.1 m, c y = 0.15 m and u = 0.35 m. Radius varies from 0.25 m to 0.5 m with an increment of 0.05 m.

P.e. Kapur, T.P. Meloy / Int. J. Miner. Process. 53 ([998) 15-28

27

7. Summary and conclusions The model presented here is at best only an exploratory foray and it undoubtedly has considerable scope for improvement and refinement. Even though there is some uncertainty regarding the attainment of a steady state in the process, the simulation results suggest that the force equilibrium approach could provide a useful theoretical frame-work for the model. The admittedly crude averaging approximation made for the flow profile is justified by the need for developing a working model in an operational sense in face of the sheer complexity of the hydrodynamic regime prevailing in the system. Even then, further refinement of the flow description, without excessive and elaborate computations, remains a challenging task for the future. No attempt was made to fine-tune the various parameters required to implement the model, that is, tailor them to a spiral operation in order to improve the model performance. In any case, this exercise cannot be undertaken without a fairly extensive and reliable data base which unfortunately is not available, at least in the public domain. Moreover, it is also necessary to account for the stochastic component in the working of the spiral which in fact is inherent in all gravity separation processes. In spite of these drawbacks, we are sufficiently encouraged by the simulation results to believe that the model presented here has considerable potentials for analyzing spiral design and its separation characteristics for different feed materials which may range from relatively light coal to heavy high-grade minerals.
References
Allen, J.R.L., 1982a. Sedimentary Structures: Their Character and Physical Basis, Vol. I. Elsevier, Amsterdam, Chapters 2 and 7. Allen, J.R.L., 1982b. Sedimentary Structures: Their Character and Physical Basis, Vol. II. Elsevier, Amsterdam, Chapters 2 and 3. Bagnold, R.A., 1954. Experiments on a gravity-free dispersion of large spheres in a Newtonian fluid under shear. Proc. R. Soc., London, Ser. A 225, 49-53. Bagnold, R.A., 1966. An approach to the sediment transport problem from general physics. U.S. Geol. Surv. Prof. Pap. 422-1. Burt, R.O., 1984. Gravity Concentration Technology. Elsevier, Amsterdam, pp. 261-287. Chang, H.H., 1988. Fluvial Processes in River Engineering. John Wiley, New York, Chapter 8. Chen, Ci-L, 1991. Power law of flow resistance in open channels: Manning'S formula revisited. In: Yen, B.C. (Ed.), Channel Flow Resistance: Centennial of Manning's Formula. Water Resources Publications, pp. 206-240. Dingman..S.L., 1984. Fluvial Hydrology. Freeman, New York, Chapter 6. Holland-Batt, A.B., 1989. Spiral separation: theory and simulation. Trans. Inst. Min. Metall. (Sect. C) 98, C46-C60. Holland-Batt, A.B., Holtham, P.N., 1991. Particle and fluid motion on spiral separators. Miner. Eng. 4, 457-482. Holtham, P.N., 1990. Flow visualisation of secondary currents on spiral separators. Miner. Eng. 5, 279-286. Holtham, P.N., 1992. Primary and secondary fluid velocities on spiral separators. Miner. Eng. 5,79-91. Kapur, PC., Meloy, T.P., 1996. Spirals unveiled, Part 1. Geometric Description and Physical Processes (in press). Li, M., Jancar, T., Holtham, P.N., Davis, J.1., Fletcher, CA., 1995. Approaches to the development of coal spiral models. In: Kawatra, S.K. (Ed.), High Efficiency Coal Preparation: An International Symposium. Soc. Mining, Metallurgy and Exploration, Littleton, Co., pp. 335-345.

28

P.e. Kapur, TP. Meloy / Int. 1. Miner. Process. 53 (1998) 15-28

Loveday, G.K., Cilliers, J.1., 1994. Fluid flow modelling on spiral concentrators. Miner. Eng. 7, 223-237. Mankosa, M.1., Stanley, F.L., Honaker, R.Q., 1995. Combining hydraulic classification and spiral concentration for improved efficiency in fine coal recovery circuits. In: Kawatra, S.K. (Ed.), High Efficiency Coal Preparation: An International Symposium. Soc. Mining, Metallurgy and Exploration, Littleton, Co., pp. 99-107. Palowitch, E.R., Deurbrouck, A.W., Parsons, T.H., 1991. Wet fine particle concentration, Section 2. Hydraulic concentration. In: Leonard, J.W. (Ed.), Coal Preparation, 5th ed. Soc. Mining, Metallurgy and Exploration, Littleton, Co., pp. 435-449. Simon, D.B., Senturk, F., 1976. Sediment Transport Technology. Water Resources Publication, Littleton, Co., Chapter 7. Sivamohan, S., Forssberg, E., 1985. Principles of spiral concentration. Inter. J. Miner. Process. 15, 173-181. Tavares, L.M., Sampaio, C.H., 1995. Spiral concentration for cleaning fines from major Brazilian coalfields. In: Kawatra, S.K. (Ed.), High Efficiency Coal Preparation: An International Symposium. Soc. Mining, Metallurgy and Exploration, Littleton, Co., pp. 129-138. von Seggern, D.H., 1990. CRC Handbook of Mathematical Curves and Surfaces. CRC Press, Boca Raton, Calif., pp. 206-209.

Вам также может понравиться