You are on page 1of 8

Available online at www.sciencedirect.

com

Chemical Engineering and Processing 47 (2008) 192–199

Vortex finder optimum length in hydrocyclone separation


Lucı́a Fernández Martı́nez ∗ , Antonio Gutiérrez Lavı́n ∗ ,
Manuel Marı́a Mahamud, Julio L. Bueno
Department of Chemical Engineering and Environmental Technology. University of Oviedo, C/Julián Claverı́a s/n, 33071 Oviedo, Spain
Received 4 August 2006; received in revised form 27 October 2006; accepted 7 March 2007
Available online 16 March 2007

Abstract
Effectiveness of hydrocyclone separations is highly dependent on their geometrical characteristics such as: chamber dimensions, aperture
diameters or feed inlet geometry, for instance. Moreover, slight modifications of any of these features might severely affect separation efficiency.
This work highlights the fundamental significance of the position of the vortex finder, showing how small changes in its length have meaningful
effects on mass recovery and particle size distribution in overflow and underflow streams. This parameter has been scarcely considered in design
studies. In order to establish the importance of the vortex finder length different and complementary methodologies were used such as mass balance,
granulometric analysis and efficiency evaluation. Results obtained using theses methodologies were in agreement, showing that the highest efficient
length of the vortex finder is 10% of the total length of the cyclone (0.1 Lt ). This result was found for two hydrocyclones of different sizes, giving
a more consistent conclusion.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Hydrocyclones; Particle analysis; Solid–liquid separation; Vortex finder

1. Introduction Generally, the feed slurry is introduced into the hydrocyclone


flowing tangentially to the cylindrical upper zone, allowing a
Cyclones are widely known in practice, mainly due to their progressive separation of the suspended solids from the feed
use in particle separation from gaseous streams [1,2]. Applying stream. The separation principle is based on inertial forces, since
this basic knowledge to the separation of suspended solids from the circular trajectory induces a radial acceleration. If the density
liquid streams the inertial devices known as “hydrocyclones” of solid particles is higher than the fluid density, these particles
appear. are moved towards the wall and leave the hydrocyclone prefer-
A hydrocyclone is able to separate or concentrate suspended entially through the lower exit. If the particles are lighter than
particles from a fluid stream. Nowadays, hydrocyclones are used the liquid, they are drawn mainly to the upper exit.
to separate solid–fluid streams [3] fluid–fluid streams [4,5] and
gas–liquid streams [6]. 2. Separation mechanism
Cyclones are inertial devices that allow separation or concen-
tration due to the difference between inertial forces that induce The predicted helicoidal flow determines both the particle
the movement of suspended solids in a liquid bulk. Unlike separation performance and the solids distribution within a
conventional centrifuges, which use a similar separation prin- hydrocyclone [9]. There exist two theories on particle separation
ciple, hydrocyclones present many advantages [7,8], such as within a hydrocyclone. The classical Eulerian one, establishes
the absence of moving parts, low energy consumption and low that the flow within a hydrocyclone is a balance between the
residence time. radial inward drag force and the outward radial centrifugal force.
On the other hand, the Lagrangian model or particle tracking
theory establishes that the separation mechanism is driven by tur-
bulent radial fluctuations and explicitly enforced force balance.
∗ Corresponding authors. Tel.: +34 985103518; fax: +34 985103434. In fact, both theories are complementary.
E-mail addresses: lusifm@gmail.com (L.F. Martı́nez), agl@uniovi.es For dilute systems where the volume occupied by particles
(A.G. Lavı́n). may be overlooked. The Eulerian–Lagrangian model can be used

0255-2701/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2007.03.003
L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199 193

as proposed by Ma et al. [10]. Other models are proposed by


Nowakowski et al. [11], if the concentration of particles exceeds
5% by volume. The particle concentration has an influence on
viscosity stresses and, if concentration rises to 10% by vol-
ume, the larger particles move towards the less shear strain zone
(towards the air-core) and the smaller particles move towards
the wall (greater shear strain). In the mentioned study, low solid
concentration fields are measured by dual-planar laser-induced
fluorescence and the high solids concentration fields by electrical
impedance tomography (EIT) [11].

3. Experimental set-up

Two hydrocyclone sizes (5 and 10 cm i.d.) have been used


to carry out the experiments described here and to contrast
data. They have been designed according to Rietema criterion
[12]. Hydrocyclone design parameters following this criterion
Fig. 2. Schematic diagram of a conventional hydrocyclone.
are calculated based on semi-empirical equations and dimen-
sionless numbers proposed by Svarovsky [13] and Castilho and
Medronho [14]. These equations can also help to predict hydro- enters tangentially to the cylindrical hydrocyclone body which
cyclone performance. The geometry of the hydrocyclones used length is defined as ℓ.
is illustrated in Figs. 1 and 2. Table 1 shows the dimensions of A scheme of the experimental apparatus is shown in Fig. 3.
the prototypes used in our study. As it is shown, the inlet pipe This consists in a tank provided with a stirrer, which maintains
is a cylindrical tube, of 9 cm in length for the 5 cm i.d. hydrocy- the particles in suspension. The closed-circuit mode of operation
clone and of 18 cm in length for the 10 cm i.d. hydrocyclone that means that the discharged underflow and overflow are returned to
the feed tank maintaining the concentration at a constant value.
Water at room temperature (15 ◦ C) is fed using a 3 kW centrifu-
gal pump P050/30T, passing through a by-pass that regulates the
flow, which is controlled by a Khrone Aquaflux 090 K/D DN40
PN 40 electromagnetic flow meter. The hydrocyclones are con-
nected in parallel. The existence of two valves allows operating
with a single hydrocyclone in all experiments.

4. Materials and methods

The feed sample was taken directly from the tank. Overflow
and underflow samples were taken from the tank return lines. A
granulometric analysis was carried out and the suspended solids
(SS) concentration was measured for all samples.
The mass of suspended solids is measured after drying a
known volume (at 105 ◦ C for a minimum of 6 h) and weight-
ing by difference. The suspended solids tested are composed by
CaCO3 with a purity of 81%. According to Perry [15], the value

Fig. 1. Some typical views (ground and side) and a schematic 3D diagram of a
hydrocyclone.

Table 1
Dimensions (centimetres) of the hydrocyclones used in the experiments
DC (internal) Di Do ℓ Du Lt − ℓ θ

5 (m) 1.4 1.7 4 1 21 11.4◦


10 (cm) 2.9 3.4 8 2 42 11.4◦
Fig. 3. Simplified flow sheet of the experimental plant.
194 L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199

of density for CaCO3 is within 2 and 2.8 g/mL. The real den-
sity of the CaCO3 used in this study (2.75 g/mL) is measured
by AccuPyc 1330 V2.04N. Particle size distribution (by volume
percentage) is measured by a laser granulometer MALVERN
MASTERSIZER QS (small volume sample dispersion unit).

5. Results and discussion

5.1. System overall balance

Laser analysis provides the volume of particles of a given


diameter. From this analysis D50 is defined [16] as the particle
diameter for which the accumulated mass is 50%, that is, in fact,
the mass median of the distribution.
Only with the determination of D50 is it not possible to obtain
the real efficiency of a hydrocyclone, which makes it necessary to
carry out other determinations or calculations in the processing
of the experimental data.
Considering a two-phase system (solid–liquid system) as
Fig. 5. Differential forms. (a) Absolute mass distribution, (b) relative fraction
schematically shown in Fig. 4, being mF the mass of fluid, mS frequency curve f(dp ), (c) cumulative fraction frequency curve F(dp ).
the total mass of solids and mSij the mass of a class of solids
whose diameter is between dpi and dpj (dpi < dp < dpj ).
Being ⌢yS provided by drying and weighting and ⌢ xij by particle
Thus, the concentration can be expressed in different ways,
size analysis.
depending on its reference basis which may be the suspension,
The characteristics of the different streams may be described
the fluid or the dried solid.
by their composition, expressed not only as total solid concen-
mS
=⌢ yS ; mass fraction of solids (1) tration but also in terms of concentration of each class (discrete
mT or differential) of solids. The distribution of solids may thus
mS ⌢ be expressed by functions which are shown in Fig. 5. The left
= Y S ; mass relation of solids (2)
mL hand-side image shows mass versus particle diameter (objective
mSij function in our case), the right hand-side figure being its nor-
=⌢ yij ; mass fraction of class ij (3) malised form, frequency versus particle diameter. Frequency
mT
may be expressed, according to the method of granulometric
mSij ⌢
= Y ij ; mass relation of class ij (4) analysis as frequency in mass or volume, in surface or in num-
mL ber. D50 , is the median, a previously mentioned form of mean
mSij diameter.
= x̂ij ; fraction of the mass of solids of class ij (5)
mS Once the concentration variables have been defined, a global
Thus, balance may be applied to the separator. Following the nomen-
clature defined in Fig. 6, let us first define the ratio between the
mSij = mS ⌢
xij = mT ⌢
yij (6) underflow and the feed flow as R1 , and the ratio between concen-

mS ⌢ tration in the underflow and feed as R2 . W is used to define mass

yij = xij = ⌢
yS ⌢
xij (7) flow, the subscript “W” being used for the underflow stream,
mT

Fig. 4. Schematic macroscopic representation of two-phase system. Fig. 6. Separator global mass balance.
L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199 195

the subscript “F” for the feed stream and “O” for the overflow
stream.
⌢ ⌢
WO WW
⌢ = 1 − R1 ; ⌢ = R1 ; 0 ≥ R1 ≤ 1 (8)
WF WF

y
R2 = ⌢SW ; ∞ ≥ R2 ≥ 1 (9)
ySF
A global balance of suspension may be written as:
ŴF = ŴO + ŴW (10)
and the solids balance as:
ŴSF = ŴSO + ŴSW (11)
⌢ ⌢ ⌢
ŴF ySF = ŴO ySO + ŴW ySW (12)
In the same way, the partial solids balance of ij class is:
Fig. 7. Elimination of short-circuit by means of a vortex finder.
ŴijF = ŴijO + ŴijW (13)

ŴF ⌢
yijF = ŴO ⌢
yijO + ŴW ⌢
yijW (14) the case of Rietema [12] who gives a value of 0.4; Bradley [17]
and Hass et al. [18] propose a figure of 1/3, Wang et Yu [19]
ŴF x̂ijF = ŴO x̂ijO + ŴW x̂ijW (15) accept as valid a value of 0.67; Narasimha et al. [20] use two
values (0.67 and 0.5). Kraipech [21] juggle with different vortex
Thus, the total separation efficiency (E) is: finder insert depths but for different hydrocyclone geometries,
⌢ ⌢
underflow solids W SW WW ⌢y obtaining ratios ranging from 0.28 to 0.93.
E= = ⌢ = ⌢ ⌢SW In this study, different vortex finder lengths have been tested
feed solids W SF W F ySF

y in order to understand the influence of this parameter on con-
= R1 ⌢SW = R1 R2 (16) centration and particle distribution, that is, their influence on
ySF
efficiency; comparing their value with those proposed by other
and the partial separation efficiency (Eij ) is: authors.
⌢ For the hydrocyclone measuring 5 cm in i.d., working with
(ij)class of solids in the underflow WijW
Eij = = ⌢ different feed flows the effect of vortex finder depth on concen-
(ij)class of solids in the feed WijF tration has been evaluated (see Fig. 8). It can be seen that trend
⌢ ⌢ ⌢ ⌢
WW yijW yijW xijW is similar for the different flows. However, the higher the flow,
= ⌢ ⌢ = R1 ⌢ = R1 R2 ⌢ (17)
WF yijW yijW xijW the more elevated concentration. According to the maximum
efficiency point, when the ratio between the vortex finder length
and total length is 0.1, the underflow concentration reaches the
5.2. Calculation of the vortex finder optimum length highest value, and thus resulting in maximum efficiency.
For further understanding of separation behaviour, the above
The insertion of the vortex finder attempts to avoid the results are compared with those obtained by granulometric anal-
re-entrainment of particles in the overflow stream [17]. This ysis.
element avoids the so-called “short-circuit” generated at the top Fig. 9 shows the particle size distribution curves for the
portion of the hydrocyclone, close to the feed inlet and the over- most efficient point, obtained when the depth of the vortex
flow upper exit (Fig. 7). Thanks to the vortex finder, the particles finder is 2.5 cm, (vortex finder depth-hydrocyclone length ratio
are induced to flow down guided by the outside wall. Increas-
ing the vortex finder length, more time is given for particle
re-entrainment in the underflow stream and this increases sepa-
ration efficiency. Nonetheless, if the vortex finder tip reaches the
conical zone, some coarse particles might reach the return over-
flow stream instead of exiting through the apex and this causes a
decrease in efficiency. Optimum length depends on feed particle
size and distribution and this should be determined preferably
by experimentation.
There does not exist a complete agreement of vortex finder
length due to the fact that this depends on geometry, feed particle
size and feed concentration. There are several values recorded in
the bibliography that express the ratio of the length of the vortex Fig. 8. Solids concentration in the underflow stream vs. fraction length of vortex
finder to the hydrocyclone diameter (Lvortexfinder − DC ). This is finder for the 5 cm i.d. hydrocyclone.
196 L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199

study shows a high turbulence zone in the region of transition


between the conical and cylindrical part that leads to the least
efficient point. This is in accordance with predictions of Dai et
al. [22] and Yang et al. [23]. When there is no introduction of
the vortex finder, some of the particles go through the upper exit
without passing into the hydrocyclone. Introducing the vortex
finder, this effect can be avoided. Nevertheless, if there is an
excessive depth, some of the particles can reach the overflow
stream instead of the underflow stream due to the swirls gener-
ated in the conical part. In our experiments, we have tried to find
the optimum point.
Fig. 10 shows that at the most efficient point (2.5 cm vortex
finder length), the vortex finder tip is situated about one centime-
tre under the lowest point of the feed inlet section. This allows
a re-entrainment of particles within the hydrocyclone, avoiding
the “short-circuit”. Nonetheless, if the vortex finder reaches the
conical section, the existing turbulence could cause coarser par-
ticles to exit through the overflow stream instead of through the
underflow stream as expected. This has also been demonstrated
by other further experiments carried out by the authors regarding
pressure fields.
By using 10 cm i.d. hydrocyclone, and following the same
experimental procedure, it has also been proved that the higher
the feed flow, the better the efficiency.
Fig. 9. Granulometric analysis for the most and least efficient configurations for For the intermediate flow (7.5 m3 /h), the mass balance also
the 5 cm i.d. hydrocyclone. shows an optimum condition (highest solids amount eliminated
by the underflow) when the ratio between the vortex finder length
of 0.1) and for the least efficient point (vortex finder depth- and the total hydrocyclone length is again 0.1, 5 cm being the
hydrocyclone length ratio of 0.2). introduced length of the vortex finder.
It can be seen that for the optimum ratio (0.1) a higher number Once the mass balance has been studied, it is necessary to
of coarser particles are recovered in the underflow stream. When test the reliability of the results obtained by laser granulometric
the ratio is 0.2 the curve corresponding to the underflow stream analysis. For the same intermediate flow, Fig. 11 shows the par-
is displaced to the right-handside and lower recovery values are ticle size distribution for the most efficient configuration (ratio
achieved. A hydrocyclone is more effective in solid separation, between the vortex finder length and the total length of 0.1) and
the higher the mass and the coarser the particles recovered by for the least efficient (ratio equal to 0.2). If these two plots are
the spigot. compared, it can be seen that the highest percentage in volume
Fig. 10 shows a scheme of the most and least efficient vortex of coarser particles in the underflow is obtained when the ratio
finder configurations. The changes in efficiency with the length is 0.1.
of the vortex finder might also be explained by the pressure tur- Hydrocyclones should preferentially separate coarser parti-
bulences within a hydrocyclone (as tested by the authors in recent cles. So, parameter D50 is used in this article to complement
experiments on pressure patterns within a hydrocyclone). This information obtained by mass balance. These results are shown
in Fig. 12 when the maximum value of D50 is obtained at the
vortex finder–total length ratio equal to 0.075 and 0.1 for the
hydrocyclones of 10 and 5 cm i.d., respectively. These values
are close to optimum ratio (0.1) obtained by mass balance.
For further analysis of what occurs within a hydrocyclone,
granulometric analysis is carried out when the ratio reaches 0.3
(Fig. 13). It can be seen that overflow and underflow streams
overlap, so there is a division but no separation. This again evi-
dences that the introduction of the vortex finder in the conical
space does not aid the separation process.
Similar results were presented by Kraipech et al. [21]. This
study shows how hydrocyclone geometry and type of feed slurry
determine the separation process. Due to this fact, only by mod-
ifying the vortex finder length, may the entire separation process
Fig. 10. Most and least efficient vortex finder lengths (left and right, respec- be modified. This is an essential characteristic of the versatility
tively) in the 5 cm hydrocyclone (dimensions given in centimetres). of hydrocyclones for different industrial purposes.
L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199 197

Fig. 12. D50 values vs. fractional length of the vortex finder for the 5 and 10 cm
i.d. hydrocyclone.

Fig. 11. Granulometric analysis for the most and least efficient configurations
for the 10 cm i.d. hydrocyclone.

Fig. 14 shows different configurations of the vortex finder


within the 10 cm i.d. hydrocyclone in order to explain the influ-
ence of the vortex finder depth on the separation process. When
the tip of the vortex finder is still placed in the cylindrical part,
this allows a re-entrainment of particles within the hydrocyclone Fig. 13. Granulometric analysis for 0.3 vortex finder length–total length ratio
avoiding “short-circuiting”. Nonetheless, if the tip of the vortex for the 10 cm i.d. hydrocyclone.
finder is placed in the juncture between the conical and cylin-
drical regions, the turbulences generated within this zone avoid For a more in depth approach to the reliability of the results
a correct separation and could cause coarser particles to exit obtained, and taking into account the mathematical model pre-
through the overflow stream instead of through the underflow viously explained, complementary analysis of the data obtained
stream as expected. Fig. 14 also shows a longer vortex finder is carried out based on hydrocyclone efficiency.
(15 cm) its tip being situated in the conical part of the hydro- Efficiency is calculated according to the Eq. (16); using the
cyclone when the ratio of this length to the total hydrocyclone parameters R1 and R2 defined in Eqs. (8) and (9), respectively,
length is 0.3. the underflow concentration and a constant value of feed con-

Fig. 14. Three vortex finder lengths in the 10 cm hydrocyclone (dimensions given in centimetres).
198 L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199

Table 2 Appendix A. Nomenclature


Calculus of efficiency of 10 cm diameter hydrocyclone (R1 = 0.098; CF = 0.8 g/L)
Lvortex finder /Ltotal Underflow R2 Efficiency (E)
concentration (g/L) CF feed concentration (kg/L)
0 3.03 3.79 0.37
dp particle size (m)
0.07 3.13 3.91 0.38 D50 particle diameter for which half of accumulated mass
0.1 3.89 4.86 0.48 achieved (m)
0.2 1.32 1.65 0.16 Dc hydrocyclone cylindrical section diameter (m)
0.3 2.32 2.90 0.28 Di hydrocyclone inlet diameter (m)
Do hydrocyclone vortex finder diameter (m)
Du hydrocyclone apex diameter (m)
centration (0.8 g/L). These values together with length ratios are E total separation efficiency
shown in Table 2. Eij partial separation efficiency
As proven by the previous analysis, the optimum lengths ratio ℓ cylindrical part length (m)
corresponds to 0.1 with efficiencies close to 50%. Lt total length (m)
Hence, an agreement in data is demonstrated using different Lvortex finder vortex finder clearance (m)
methodologies such as mass balance, calculus of efficiency and mL mass of fluid (kg)
granulometric analysis. This allows us to carry out an in-depth mS total mass of solids (kg)
study of hydrocyclone behaviour, finding a value of optimum mSij fraction of a class of solids whose diameter is
vortex finder depth between those mentioned in the biblio- dpi < dp < dpj (kg)
graphy. mT total mass (kg)
R1 underflow-feed mass flow ratio
R2 underflow-feed concentration ratio
6. Conclusions
x̂ij fraction of the mass of solids of class ij

yS mass fraction of solids
The following conclusions can be drawn from this study: ⌢
yij mass fraction of class ij

YS mass relation of solids

• The depth at which the vortex finder tip is placed greatly Y ij mass relation of class ij
influences hydrocyclone efficiency. ŴO overflow mass flow (kg/s)
• This optimum vortex finder length value corresponds to a ŴijO overflow mass flow of class ij (kg/s)
vortex finder length-hydrocyclone total length ratio of 0.1. ŴF feed mass flow (kg/s)
This optimum value is obtained by the analysis of both the ŴijF feed mass flow of class ij (kg/s)
solid mass separation in the underflow stream. Moreover, this
ŴW underflow mass flow (kg/s)
value is in the range of those presented in the literature.
ŴijW underflow mass flow of class ij (kg/s)
• For hydrocyclones designed considering geometrical simi-
larity (proportional dimensions), the optimum ratio of vortex References
finder length to total hydrocyclone length tends to be the same,
regardless of the hydrocyclone size and thus differences in [1] M. Crawford, Air Pollution Control Theory, McGraw-Hill, NY, 1976.
values of concentration and D50 . [2] D.F. Ciliberti, B.W. Lancaster, An improvement of the simple model for
• In the absence of vortex finder, the short-circuit generated in rotatory flow cyclones, AIChE J. 22 (1976) 1150–1152.
[3] M.J. Doby, W. Kraipech, A.F. Nowakowski, Numerical prediction of out-
the upper part of the hydrocyclone avoids a clear separation. In
let velocity patterns in solid–liquid separators, Chem. Eng. J. 111 (2005)
the same way, when the depth of the vortex finder tip is exces- 173–180.
sive, a substantial decrease in efficiency may be observed due [4] C.A. Petty And, S.M. Parks, Flow structures within miniature hydrocy-
to the swirls generated at the bottom of the hydrocyclone. clones, Miner. Eng. 17 (2004) 615–624.
• Less efficient conditions were found when the vortex finder [5] R. Delfos, S. Murphy, D. Stanbridge, Z. Olujic, P.J. Jansens, A design tool
for optimising axial liquid–liquid hydrocyclones, Miner. Eng. 17 (2004)
depth is near the juncture between the cylindrical and the
721–731.
conical part, where a high turbulence may appear due to the [6] J.R. Parga, D.L. Cocke, Oxidation of cyanide in a hydrocyclone reactor by
synergy of two phenomena: the change of trajectory by enter- chlorine dioxide, Desalination 140 (2001) 289–296.
ing in the conical part and the turbulence associated to the [7] K. Dwari, M.N. Biswas, B.C. Meikap, Performance characteristics for par-
vortex finder itself. This is in accordance with predictions of ticles of sand FCC and fly ash in a novel hydrocyclone, Chem. Eng. Sci.
59 (2004) 671–684.
Dai et al. [22] and Yang et al. [23].
[8] K. Udaya Bhaskar, B. Govindarajan, J.P. Barnwal, K.K. Rao, B.K. Gupta,
T.C. Rao, Classification studies of lead–zinc ore fines using water-injection
Acknowledgments cyclone, Int. J. Miner. Process. 77 (2005) 80–94.
[9] L. Fernández, A.G. Lavı́n, M.M. Mahamud, J.L. Bueno, Hydrocyclones in
water depuration (Spain), Tecnologı́a del Agua 275 (2006) 42–56.
The authors would like to acknowledge the financial support [10] L. Ma, D.B. Ingham, X. Wen, Numerical modelling of the fluid and particle
from the “Ministerio de Educación y Ciencia” (Spain) for this penetration through small sampling cyclones, J. Aerosol. Sci. 31 (2000)
investigation within the Project (REN2003-09389). 1097–1119.
L.F. Martı́nez et al. / Chemical Engineering and Processing 47 (2008) 192–199 199

[11] A.F. Nowakowski, J.C. Cullivan, R.A. Williams, T. Dyakowski, Appli- [17] D. Bradley, The Hydrocyclone, Pergamon Press, Oxford, 1965.
cation of CFD to modelling of the flow in hydrocyclones. Is this a [18] P.A. Haas, E.O. Nurmi, M.E. Whatley, J.R. Engel, Midget hydroclones
realizable option or still a research challenge? Miner. Eng. 17 (2004) 661– remove micron particles, Chem. Eng. Progr. 53 (1957) 203–207.
669. [19] B. Wang, A.B. Yu, Numerical study of particle-fluid flow in hydrocyclones
[12] K. Rietema, Performance and design of hydrocyclones, Parts I–IV, Chem. with different body dimensions, Miner. Eng. 19 (2006) 1022–1033.
Eng. Sci. 15 (1961) 298–325. [20] M. Narasimha, M. Brennan, P.N. Holtham, Large eddy simulation of
[13] L. Svarovsky, Solid-Liquid Separation, fourth ed., Butterworth- hydrocyclone-prediction of air-core diameter and shape, Int. J. Miner.
Heinemann, Oxford, UK, 2000. Process. 80 (2006) 1–14.
[14] I.R. Castilho, R.A. Medronho, A simple procedure for design and perfor- [21] W. Kraipech, W. Chen, F.J. Parma, T. Dyakowski, Modelling the fish-hook
mance prediction of Bradley and Rietema Hydrocyclones, Miner. Eng. 13 effect of the flow within hydrocyclones, Int. J. Miner. Process. 66 (2002)
(2000) 183–191. 49–65.
[15] R.H. Perry, Perry’s Chemical Engineers’ Handbook, seventh ed., McGraw- [22] G.Q. Dai, J.M. Li, W.M. Chen, Numerical prediction of the liquid flow
Hill, NY, 1999. within a hydrocyclone, Chem. Eng. J. 74 (1999) 217–223.
[16] C. Puprasert, G. Hebrard, L. Lopez, Y. Aurelle, Potential of using hydrocy- [23] I.H. Yang, C.B. Shin, T.-H. Kim, S. Kim, A three-dimensional simulation
clone and hydrocyclone equipped with grit pot as a pre-treatment in run-off of a hydrocyclone for the sludge separation in water purifying plants and
water treatment, Chem. Eng. Process. 43 (2004) 67–83. comparison with experimental data, Miner. Eng. 17 (2004) 637–641.

You might also like