Sie sind auf Seite 1von 11

Theoretical Foundations of Chemical Engineering, Vol. 34, No. 5, 2000, pp. 428--438.

Transkltedfrom Teoreticheskie Osnovy Khimicheskoi Tekhnologii, Vol. 34, No. 5, 2000, pp. 478--488. Original Russian Text Copyright 9 2000 by Dueck, Matvienko, Neesse.

Modeling of Hydrodynamics and Separation in a Hydrocyclone


J. G. Dueck*, O. V. Matvienko**, and T. Neesse*
* Friedrich-Alexander Universititt Erlangen-Niirnberg, LUR, Schottkystr. 10, 91058 Erlangen, Germany ** Tomsk State University of Architecture and Construction, Solyanaya pl. 2, Tomsk, 634003 Russia ReceivedAugust I!, 1999

Abstract--A numerical study of the flow pattern and the separation processes in a hydrocyclone was performed on the basis of the Navier-Stokes equations. The mathematical model presented here enabled us to calculate the separation characteristics and the distributions of the particle velocity, pressure, and particle concentration in the apparatus. Hydrocyclones are used for separating the solid material from a suspension and also for its classification (separation into fractions). These apparatuses are comparatively simple in design, compact, and economical. The widest industrial application has been found by tapered cylindrical hydrocyclones (Fig. 1). The theoretical foundations of the hydromechanics of separation in hydrocyclones and related separation processes have been developed for many decades [1]. Initially, the researchers restricted themselves to deriving empirical formulas for the separation characteristics of a hydrocyclone from experimental data [2]; or, if the information on the flow pattern in the apparatus was more or less reliable (see, e.g., [3]), attempts were made to derive relatively simple analytical expressions. As a rule, the deterministic motion of single particles was considered [4, 5]. The works along this line of research have revealed a number of peculiarities of the particle motion in a hydrocyclone and have enabled one to make a general assessment of the effect of some parameters of the apparatus and of the mixture on the separation characteristics. However, the failure to take account of the influence of random forces (which inevitably arise under turbulent conditions and give rise to diffusion flow of particles) and also the neglect of the reverse effect of the solid phase on the flow necessitate substantial empirical correction of the deduced formulas. In [6-9], a model of separation of the solid phase in the apparatus has been put forward. This model is based on the notion of the turbulent diffusion of particles, which induces a solid-particle flow whose direction is opposite to that of the sedimentation flow caused by centrifugal forces. In those works, the concentration field and the separation characteristics have also been calculated. Such calculations in a closed analytical form are possible only after the actual hydrodynamic situation in the apparatus is essentially simplified namely, the turbulence is assumed to be uniform throughout the apparatus and a fairly simple model of the real hydrodynamic field is adopted. Obviously, this oversimplification prevents this separation model from being applied in practice without invoking additional empirical relationships. The modern methods of hydrocyclone design are based on computational mathematics. Their success depends primarily on the progress in modeling of highly swirling turbulent multiphase flow. Many details of this modeling still remain to be refined. However, this does not detract from (but rather enhances) the significance of the attempts to use the complete hydromechanic equations (the Navier-Stokes equations), the equations of fairly developed turbulent flow, and the equations of phase interaction mechanics in the design of relevant apparatuses. Such calculations can be important in practice (this point is corroborated by the

Oh

Ll

L2

t L3
Dbot
Fig. 1. Schematic of a hydrocyclone: Dh = 40 mm, DO= 12.8 mm, Dbot = 5 mm, Dtop= 16 mm, L 1 = 75 mm, L2 = 390 mm,/-3 -- 50 mm, II = 40 mm, 12= 30 mm, l = 2 ram, and 0t = 2.6~

0040-5795/00/3405-0428525.00 9 2000 MA1K"Nauka/Interperiodica"

MODELING OF HYDRODYNAMICS

AND SEPARATION IN A HYDROCYCLONE

429

present work); furthermore, they stimulate the development of related areas of continuum mechanics. Computer-oriented mathematical models have been presented in [10, 11], where many important aspects of fluid mechanics and the motion of single particles in an apparatus have been discussed. These works [10, 11] have taken no account of the turbulent diffusion of particles. However, disregarding this process may sometimes make the description of separation processes unsatisfactory. Thus, the main purpose of this work, which continues earlier studies [12, 13], is to describe the swirling turbulent flow of a suspension in relation to the interaction between the phases, and also to design the process of separation of the components. MATHEMATICAL MODEL OF HYDROCYCLONE HYDRODYNAMICS To model hydrocyclone hydrodynamics, we will use the so-called algebraic slip velocity method, which was proposed in [11]. This method is based on the equation of conservation of the momentum of a suspension, the equation of continuity (conservation of mass) of a suspension, the equation for the relative slip velocity of particles of each of the fractions, and the material balance equations for each of the fractions. The equations of conservation of mass and momentum are written as div(pzVz) = O,
(V m 9V)Vmp m =

model [14]:

div(pmVmk- lxeffgrad(k)) = ( G k - pze),


div (pmVme -- ~_-ff%ffgrad(e)~
oe ./

(4)

(5)

Gk=lX, 2 ~
+

{ E(u)

+~,~r) +

(~u + ~xr') 2 f~w'~ 2 (r~wlr)2

~,~ r

~x ) + ~-~x) + ~, "-~r J J"

Here, x and r are the axial and radial coordinates; and u, v, and w are the axial, radial, and tangential components of the velocity vector V m. The values of the constants are the following: C~t= 0.09; C1 = 1.44; C2 = 1.92; C 3 = 0.001; and o~ = 1.3 The suspension density Pz is expressed in terms of the concentrations mj of the fractions, the particle densities pj in the fractions, and the liquid density:
N

1- Zmj
i=lpl

+j~l (mj)'pj

(6)

(1) (2)

- g r a d ( p ) + div(Tz) + F z,

The dependence of the suspension viscosity on the solid-phase concentration can be taken into account by, e.g., the Thomas formula [15]
N

where V s is the suspension velocity vector with the components (v-i, v-2, and v-a); Ps is the mixture density; p is the pressure; Tm is the stress tensor; and F m are the body forces. Let us assume that the solid-phase concentration in the suspension, governing its viscosity and density, influences the velocity field of the flowing medium. Let us also assume that the suspension obeys the Newtonian rheology. The relationship between the components of the stress tensor and velocity vector will then appear as

kt__~m 1 + 2.5 Z = ~! j = 1 PJ N 2 N + 0.0275 ~ " (m exp(16.6 ~-'m .P~]

mjPm

.Pm]

(7)

)=

(3) where I.tefeis the effective viscosity, which is the sum of the viscosity I.tmof the suspension and the turbulent viscosity l.tt = Cttpmk2/e. The turbulent energy k and the dissipation rate e are computed by the equations of the modified k-e

On the right side of (7), the first two terms constitute a familiar Einstein formula and the last term is a correction that is important at high solid-phase concentrations. Formula (7) is deduced for the case when the particle densities in different fractions are unequal. After the Reynolds averaging, the equation of conservation of mass of the disperse system in the turbulent flow takes the form of a diffusion equation and can be written as

div(pmmjVrj) = -div(pmmjVm- Dtjgrad(mj)), (8) where Vrj is the relative velocity of the dispersed phase.
As in [16], the coefficient D o of turbulent diffusion of particles in (8) is calculated with allowance made for the stochastic motion of the largest turbulent vortices only. A particle from the jth fraction is entrained by a
Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

430

DUECK et al.

Fig. 2. Distributio0s of the turbulentdiffusioncoefficient((kg s)/m)of particles with the diameterd = (a) 1 and (b) 17.2/amover the hydrocycloneat Cm= 100 g/l.

turbulent vortex and moves at some frequency toj and velocity amplitude uj. Therefore, the diffusion coefficient is expressed as D o = u~ /03j. To calculate the velocity of a particle in the turbulent field, we use the equation taking into account the particle inertia:

sented as

uj

w~ exp(iCOot)exp(-iarctan(OOoXj)). 2 2 J1 + tgo'Cj

(10)

duj _ u j - w ' dt xj

(9)

Apparently, the frequency of oscillations of the velocity of particles from thejth fraction coincides with the liquid velocity oscillation frequency: coj = co0. The phase difference is immaterial. The particle oscillation amplitude is given by the first factor in (10). Taking into account the square of this amplitude leads to the following expression for the diffusion coefficient:
2

where xj is the relaxation time of the particle velocity to the liquid velocity. An unsteady (oscillating) liquid velocity can be given by the Fourier series, from which we retain only the fundamental harmonic w' = woexp(itOot). In the turbulent flow mode, the initial condition is unimportant, and the solution of Eq. (9) can be repre-

Dt j =

Wo 1 2 2 " 1 + O~oZ ~ j

(11)

For minute particles, the diffusion coefficient should 2 be equal to the turbulent viscosity: wolo) o = ~t. The fundamental frequency can be expressed in terms of turbulence characteristics: coo = e/k. The relaxation time Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONSOF CHEMICAL ENGINEERING

MODELING OF HYDRODYNAMICS AND SEPARATION IN A HYDROCYCLONE

431

for sufficiently small particles (in the Stokes flow) can be found as xj = p~(dj)2/l.tt. Under these conditions, the diffusion coefficient is DO = I+G

()2
e

Thus, there is no need to solve the complete differential equations of motion, and it is sufficient to consider the equation of the dynamic balance of forces, whose expanded form is

(12)

2(j)

~CD(dj)qlVrjlVrj + ( p j - pm)aj = 0.

(15)

~r (P J) 7 Equation (15) enables one to determine the relative dispersed-phase velocity 4 . , ,2 ( P j - Pm) Vr~ = ~ta) ~-C~-rerja~. (16)

where C, = 104 is a correction factor. Formula (12) fits the experimental data [16] well enough. Since the diffusion coefficient D o depends on the turbulence field, it changes from point to point (contrary to the simplest hypothesis accepted in [3, 6-8]). Figure 2 presents some examples of the calculated distribution of the turbulent diffusion coefficient of particles in a hydrocyclone. For small particles, the diffusion coefficient is great enough and its distribution over the apparatus is similar to the turbulent viscosity distribution (Fig. 2a). Large particles cannot diffuse, although turbulence is highly developed (Fig. 2b). The transfer of the momentum of the disperse phases is described in terms of the theory of multivelocity continuum by the Eq. [17]
( V j . V ) V jmjp m =

The drag coefficient Co is a function of the relative Reynolds number

Rerj = PtlVrjld/Pt
and is given by the expression [18] 24 3.73 -4.83 x 10-3,fR--~,j
1 + 3 x 10-6Rerj

C~

~erj + ~

+ 0.49. (17)

-mjgrad(p ) - div( Tpq)

+ FDj + Foj + Fsj.


In (13),

(13)

Faj = %pjaj
is the body force, and

aj = {g, w 2r , - v j w j / r } /
is the disperse-phase acceleration (with the gravitational, centrifugal, and Coriolis components). Combination of equations (2) and (13) enables one to find the velocity of the dispersed phase relative to the suspension:

(Vrj" V)Vrjmjpm + (Vmj" V)VrjmjPm + (Vrj" V)Vmjmjp m = m2(pj- pm)aj (14) - div(Tm) - div(Tpq) + Fsj + Foj. The turbulent stress tensor Tp# is defined similarly to (3).
Analysis of Eq. (14) suggests the following. The ratio of the convective and diffusion terms to the drag force is a small quantity on the order of (Retdj/Dh) 2 and the ratio of the lifting force to the drag force is a small quantity on the order of ,r whereas the buoyant force is of the same order as the drag force. Here, Re t = pllVt]Dh[[t l is the Reynolds number expressed in terms of the liquid parameters and the hydrocyclone diameter.

The boundary conditions at the hydrocyclone inlet are based on a number of assumptions which allow one to obviate three-dimensional calculations. Instead of considering the actual inlet flow injected tangentially from the feeding tube along the apparatus wall, the suspension is assumed to be fed through the entire top section of the hydrocyclone. The radial velocity is chosen so that the total flux through the wall corresponds to a given flow rate through the inlet pipe. The elliptic type of the equations requires one to specify the conditions at the outlets. This task is not trivial within the framework of the problem to be solved. Strictly speaking, the flow conditions at the hydrocyclone outlet should be matched with the flow conditions beyond the apparatus, e.g., in the outlet tubes. In particular, if the effluent suspension continues to rotate at the outlet, then near the hydrocyclone axis, a region of low pressure is produced, which can cause air inflow into the apparatus with the formation of an air column. Such a flow with the freely forming surface has not hitherto been described. The boundary conditions specified in this work rule out any contact of the effluent liquid with air in the immediate vicinity of the outlet. At the bottom outlet, the environment pressure is preset; at the top outlet, the flow is supposed to be at radial equilibrium:

~p Wm 2-'7 = PmT"
RESULTS AND DISCUSSION Based on the above model, we numerically studied the flow pattern and the concentration fields of the solid phase in hydrocyclone. Figure 1 shows the hydrocyclone parameters used in the calculations. The experimental data were taken from [12]. All the main calculaVol. 34 No. 5 2000

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

432 d25, ds0, d75, lam

DUECK

et al.

l"
I
93

.J
9 /"
9 ~

d25/d75

50

o.8, ~
0.4
I I I

25
9

C in X

10-2, g/l 4

4 6 Cin x 10 -2, g/l

Fig. 3. Calculated (lines) and observed .(points) dependences of (1) d25, (2) dso, and (3) d75 on C m.

Fig. 4. Calculated (line) and observed (points) dependences of d251d75 on C'n. C~ %

foul

x 10-2, g/1
9 ~

[] 2
A3
9

i
r ~

"t.f/
Jo
9 i J oz i

01

6 Cin x 10-2, g/l

10-t

10~

101

102 d, ktm

Fig. 5. Calculated (lines) and observed (points) dependences of the panicle concentrations at the (1) bottom and (2) top outlets on C m.

Fig. 6. Cal~:ulated (lines) and observed (points) separation curves at C m = (1) 100, (2) 200, and (3) 600 g/I.

tions, as well as the experiments, were carded out at the pressure 0.15 MPa. The basic quantitative parameter of hydrocyclone is the flow rate. There are many empirical relationships between the capacity and the inlet pressure [2]. The results of numerical calculations, which fit the empirical formulas well, demonstrate that the liquid flow rate is proportional to the square root of the pressure. A parameter characterizing the separation of the solid material in hydrocyclone is the so-called separation grain diameter ds0, which is the diameter of a par-

ticle that reaches the bottom outlet with the probability of 50%. Along with this quantity, the d25 and d75 values are also important, which are the sizes of grains that leave the apparatus through the bottom outlet with the probabilities 25 and 75%, respectively. These parameters are often used as hydrocyclone characteristics. The calculated and experimental data in Fig. 3 imply the explicit dependence of d25, ds0, and (/75 on the properties of the suspension, in particular, on the solid-phase concentration, ds0 can be considered constant only for highly diluted suspensions. The separation grain size can grow with increasing suspension density.
Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

MODELING OF HYDRODYNAMICS AND SEPARATION IN A HYDROCYCLONE

433

Fig. 7. Distributionsof the turbulentenergy(m2/s2) in the hydrocycloneat Cin = (a) 0 and (b) 532 g/l.

The degree of suspension separation is commonly characterized by the ratio d2Jd75. Figure 4 collates the calculated and experimental separation factors as functions of the solid-phase concentration at the hydrocyclone inlet. The d25/d75 ratio is seen to lower with increasing concentration of the injected suspension. The same trend is shown by the variation of the concentrations of the solid component in the suspensions withdrawn through the top and bottom outlets with the concentration of the introduced suspension (Fig. 5). It is of interest that, at low inlet concentrations, the composition of the suspension at the bottom outlet is more sensitive to the variation of the inlet concentration than the suspension composition at the top outlet. At

high inlet suspension concentrations, the situation is the opposite. With rising Cn, the properties of the suspensions withdrawn through the bottom and top outlets become closer to each other. This means that, in the limiting case, the apparatus performs no classifying. At the limiting, extremely high concentrations of the fed suspension, the influent flow is merely redistributed and leaves the hydrocyclone through the two outlets. By calculating the fluxes of each of the particle fractions through the top and bottom outlets, one can obtain the separation curve (percent of particles leaving the apparatus as a function of the particle size). Figure 6 displays the results of these calculations at various C ~ along with the observed data. By and large, the agreement between the calculated and experimental data can Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONSOF CHEMICAL ENGINEERING

434

DUECK et al.

Fig. 8. Distributionsof the turbulent viscosity [(kg s)/m] in the hydrocycloneat Cin = (a) 0 and (b) 532 g/I.

be considered satisfactory, in view of inevitable errors in processing grain-size data. The reported results enabled us to compute the hydrodynamic and concentration fields in a hydrocyclone for various operating conditions (Figs. 7-10). The data obtained cannot still be corroborated by experimental data; in a sense, they are predictions that seem to be useful for deepening the comprehension of the processes in a hydrocyclone. Figure 7 shows the turbulent energy distributions over the apparatus filled with pure water and with a suspension containing 532 g/1 of the solid-phase. The turbulence is distributed nonuniformly over the hydrocyclone; it is mainly concentrated near the top outlet and at the upper part of the conical section. This distribution

depends on the loading. The suspension exerts a general damping effect on the turbulence, and the turbulent energy peaks not near the vortex tube but near the joint of the cylindrical and conical sections of the apparatus. The turbulence energy is low (especially at high initial concentrations of the solid phase in the suspension) in the lower part of the conical section and at the bottom outlet. The turbulent energy largely governs the turbulent viscosity. As a consequence, the highest I.tt values are observed near the vortex catcher, under the cover of the apparatus, and near the joint of its cylindrical and conical sections (Fig. 8). It is of interest that, when a highconcentration suspension is injected, a region with a reduced turbulent viscosity appears near the axis of the Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

MODELING OF HYDRODYNAMICS AND SEPARATION IN A HYDROCYCLONE

435

Fig. 9. Distributionsof the particle concentration (g/l) in the hydrocyclone a t C in -- 228 g/1 and d= (a) 0.4, (b) 1.0, (c) 5.5, (d) 17.2, (e) 30.6, and (f) 55.5 lam.

top outlet and a zone with a high viscosity emerges in the axial zone of the lower part of the conical section. Such a turbulent viscosity distribution defines the distribution of the diffusion coefficient, at least that of small particles, as shown in Fig. 2.

The turbulent diffusion, counteracting the centrifugal forces emerging in the swirling flow, interfere with the separation. For example, small particles are spread by diffusion throughout the apparatus and are partially carried away through the bottom outlet. Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

436

DUECK et al.

Fig. 10. Distributionsof the particle concentration(g/l) in the hydrocycloneat d = 55.5 ~m and Cin = (a) 10.2, (b) 106, (c) 228, (d) 369, (e) 532, and (f) 726 g/I.

Figure 9 demonstrates the distributions of the concentrations of particles from various fractions. Of interest is Fig. 9b, which depicts the distribution for the 1-~m fraction. This fraction accumulates under the cover and near the top outlet. This intriguing phenomenon is related to the displacement of the liquid by the particles thrown onto the external wall. As already mentioned, smaller particles are spread by diffusion and larger ones move toward the external wall. Thus, particle motion against the centrifugal force is observed only for a close-cut fraction. It is still unclear how this phenomenon can influence the run of the separation curve. Large particles gather at the external wall. The

larger the particles the farther downstream they are collected. A high suspension concentration can cause bottom-outlet clogging, which does frequently happen in practice. Figure 10 displays the concentration distributions of quite large (d = 55.5 ~tm) particles over a hydrocyclone at various concentrations of the solid phase in the injected suspension. For a low-concentration suspension, the bottom outlet is virtually free, whereas for a high-concentration suspension, it is well choked. It is intriguing that calculations predict a relative rarefaction at the bottom outlet at extremely high inlet concentrations compare the distributions in Figs. 10e and 10f. Vol. 34 No. 5 2000

THEORETICAL FOUNDATIONS OF CHEMICAL ENGINEERING

MODELING OF HYDRODYNAMICS AND SEPARATION IN A HYDROCYCLONE Note, however, that calculations for such high concentrations of the solid phase in the suspension may be beyond the scope of the model, and their results should be dealt with certain care. ACKNOWLEDGMENTS This work was supported by the Alexander von Humboldt Foundation (Germany). NOTATION af---acceleration of a particle of the jth fraction in the field of the body forces, m/s2; Co---drag coefficient; C---concentration of the solid phase in the suspension, g/l; dT--diameter of the particles of the jth fraction, m; Dh--hydrocyclone diameter, m; D0---coefficient of the turbulent diffusion of particles of the jth fraction, (kg s)/m; Fo----drag force, N; Fa--body force, N; Fs---lifting force, N; Gk---dissipative function, (kg sa)/m; g--acceleration of gravity, m/s2; k--turbulent energy, m2/s2; mj---concentration of the particles ofthejth fraction, g/l; N--number of fractions; p---pressure, Pa; R--hydrocyclone radius, m; r----distance to the hydrocyclone axis, m; t--time, s; uF---particle velocity amplitude, m/s; Vm--suspension velocity vector, m/s; Vo---relative velocity of the disperse phase of thejth fraction, m/s; xs--radial velocity of the particles of the jth fraction, m/s; w~--tangential velocity of the suspension, m/s; wT--tangential velocity of the particles of the jth fraction, m/s; w'---oscillating liquid velocity, m/s; w0--oscillating liquid velocity amplitude, m/s; e---dissipation rate, m2/s3; lxr--molecular viscosity of the liquid (water), (kg s)/m; ~tefr---effectiveviscosity, (kg s)/m; I.tm--suspension viscosity, (kg s)/m; l&--turbulent viscosity, (kg s)/m; pro--suspension density, kg/m3;

437

pT---density of the particles of the jth fraction, kg/m3; xj---time of relaxation of the velocity of the particles of the jth fraction to the liquid velocity, s; toj-----particlevelocity oscillation frequency, s-t; co0--1iquid velocity oscillation frequency, s-~; R%---relative Reynolds number; Rer---Reynolds number. SUBSCRIPTS AND SUPERSCRIPTS in--hydrocyclone inlet; out--hydrocyclone outlet; j--fraction no.; /--liquid; m--suspension. REFERENCES 1. Svarovsky, L., Hydrocyclones, London: Technomic, 1984. 2. Povarov, A.I., Gidrotsiklony na obogatitel'nykh fabrikakh (Hydrocyclones at Concentrating Mills), Moscow: Nedra, 1978. 3. Ivanov, A.A., Ruzanov, S.R., and Lunyushkina, I.A., Fluid Dynamics and Separation in a Hydrocyclone, Zh. Prikl. Khim. (Leningrad), 1987, vol. 60, no. 5, p. 1047. 4. Baranov,D.A., Kutepov,A.M., and Lagutkin, M.G., Calculation of the Separation Process in Hydrocyclones, Teor. Osn. Khim. Tekhnol., 1996, vol. 30, no. 2, p. 117. 5. Kutepov, A.M., Lagutkin, M.G., and Baranov, D.A., A Method for Calculating the Separation Factors in Hydrocyclones, Teor. Osn. Khim. Tekhnol., 1994, vol. 28, no. 3, p. 207. 6. NeeBe, T. and Schubert, H., Modellierung und verfahrenstechnische Dimensionierung der turbulenten Querstromklassierung. Teil I, Chem. Tech. (Leipzig), 1975, vol. 27, no. 9, p. 529. 7. Neel3e,T. and Schubert, H., Modellierung und verfahrenstechnische Dimensionierung der turbulenten Querstromklassierung. Teil II, Chem. Tech. (Leipzig), 1976, vol. 28, no. 2, p. 80. 8. Neel3e,T. and Schubert, H., Modellierung und verfahrenstechnische Dimensionierung der turbulenten Querstromklassierung. Teil III, Chem. Tech. (Leipzig), 1976, vol. 28, no. 5, p. 273. 9. Schubert, H. and Neesse, T., A Hydrocyclone Separation Model in Consideration of the Turbulent Multi-Phase Flow, Proc. Int. Conf. on Hydrocyclones, Cambridge, 1980, p. 23. 10. Monredon, T.C., Hsien, K.T., and Rajamani, R.K., Fluid Flow Model of the Hydrocyclone: An Investigation of Device Dimensions, Int. s Miner. Process., 1992, vol. 35, no. 1, p. 65. 11. Pericleous, K.A. and Rhodes, N., The Hydrocyclone Classifier A Numerical Approach, Int. J. Miner Process., 1986, vol. 17, no. 1, p. 23. No. 5 2000

THEORETICAL FOUNDATIONSOF CHEMICALENGINEERING Vol. 34

438

DUECK et al. 16. Dueck, J. and Neesse, T., Contribution to the Analysis of Energy Spectrum and Transport Phenomena in a Turbulent Two-Phase Flow, Progress in Fluid Flow Research: Turbulence and Applied MHD. Progress in Astronautics and Aeronautics, Virginia, 1998, vol. 182, p. 173. 17. Nigmatulin, R.I., Dinamika mnogofaznykh sred (Dynamics of Muitiphase Media), Moscow: Nauka, 1987. 18. Schubert, H., Heidenreich, E., Liepe, E, and Nee6e, T., Mechanische Verfahrenstechnik, Leipzig: DVG, 1990.

12. Gerhart, C., Matvienko, O., Diick, J., and Neel3e, T., Numerische Berechnungen der Dichtstromtrennung im Hydrozyklon StrOmungen in der Verfahrenstechnik, Chemnitz: Tech. Univ. Chemnitz, 1998. 13. Dueck, J., Matvienko, O.V., and Neesse, T., Hydrodynamics and Particle Separation in the Hydrocyclone, Proc. H Int. Symp. on Two-Phase Flow Modeling and Experimentation, Pisa, 1999, vol. 2, p. 923. 14. Boysan, F., Selected Topics in Two-Phase Flow, Lectures Series, Trondheim, Norway, 1984, vol. 9, p. 137. 15. Thomas, D.G., Transport Characteristics of Suspensions, J. Colloid Sci., ! 965, vol. 20, p. 267.

THEORETICAL FOUNDATIONSOF CHEMICAL ENGINEERING

Vol. 34

No. 5

2000

Das könnte Ihnen auch gefallen