Anda di halaman 1dari 19

This article was downloaded by: [University of Otago]

On: 06 October 2014, At: 07:24


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Petroleum Science and Technology


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lpet20

Hydrocyclones for De-oiling


Applications—A Review
a a b
N. Kharoua , L. Khezzar & Z. Nemouchi
a
Department of Mechanical Engineering , Petroleum Institute , Abu
Dhabi, United Arab Emirates
b
Département de Génie Mécanique, Faculté des Sciences de
l'Ingénieur , Université Mentouri Constantine , Algeria
Published online: 06 Apr 2010.

To cite this article: N. Kharoua , L. Khezzar & Z. Nemouchi (2010) Hydrocyclones for De-
oiling Applications—A Review, Petroleum Science and Technology, 28:7, 738-755, DOI:
10.1080/10916460902804721

To link to this article: http://dx.doi.org/10.1080/10916460902804721

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Petroleum Science and Technology, 28:738–755, 2010
Copyright © Taylor & Francis Group, LLC
ISSN: 1091-6466 print/1532-2459 online
DOI: 10.1080/10916460902804721

Hydrocyclones for De-oiling Applications—


A Review

N. KHAROUA,1 L. KHEZZAR,1 AND Z. NEMOUCHI2


1
Department of Mechanical Engineering, Petroleum Institute, Abu Dhabi,
United Arab Emirates
2
Département de Génie Mécanique, Faculté des Sciences de l’Ingénieur,
Université Mentouri Constantine, Algeria
Downloaded by [University of Otago] at 07:24 06 October 2014

Abstract The de-oiling hydrocyclone is a device for liquid–liquid separation and


production water cleanup. Significant progress has been made in the development
of such a device since its first introduction and use in the late 1970s. The present
article is a literature review of development and research work performed so far on
de-oiling hydrocyclones. It reviews the performance parameters affecting the de-oiling
hydrocyclone operation; namely, the inlet oil concentration and drop size distribution,
the turn-down ratio, the pressure drop ratio, the flow split, and the geometrical
parameters. The article addresses work done to elucidate the internal flow structure
and performance of de-oiling hydrocylones in both experimental and computational
fluid dynamics areas. Challenges and remaining research issues are also identified.

Keywords de-oiling hydrocyclone, fluid mechanics of hydrocyclones, liquid–liquid


separation, multiphase flows, performance parameters, turbulence

Introduction
Cyclones are simple devices used to separate a dispersed phase from a continuous phase
based on centrifugal force. When the primary phase is a liquid, the device is called a
hydrocyclone. It was invented in the late 18th century (Bhaskar et al., 2007) and was
designed initially for solid–liquid separation. A large amount of literature already exists
on solid–fluid hydrocyclones and is not the focus of this article, which is concerned with
liquid–liquid hydrocyclones in de-oiling applications where the oil exists as a dispersed
phase within water as a background phase.
The most recent and challenging application of hydrocyclones to realize is liquid–
liquid separation because of the very low density differential between the phases to sepa-
rate and also because of the presence of conflicting complex flow phenomena like droplet
breakup and coalescence. Among liquid–liquid applications, the de-oiling hydrocyclone
has become the best substitute to the conventional bulky installations used to separate
fine oil droplets from produced water in the oil industry and especially so when space is
at a premium either on offshore platforms or in downhole locations. The hydrocyclone
belongs to the class of mechanical separation technologies that involves sedimentation,
centrifugation, flotation, filtration, and membrane systems. These techniques are found in

Address correspondence to Dr. Lyes Khezzar, Department of Mechanical Engineering,


Petroleum Institute, Abu Dhabi, P.O. Box 2533, United Arab Emirates. E-mail: lkhezzar@pi.ac.ae

738
Hydrocyclones for De-oiling Applications 739

various industries such as mining, upstream and downstream oil and gas, petrochemical,
and water. The development of the de-oiling hydrocyclone has progressed hand in hand
with the emergence of environmental legislation requirements for disposal or reuse of
treated produced water in on- and off-shore oil fields.
The period of the 1970s and the 1980s saw tremendous achievements with regard
to liquid–liquid separation. The main thrust was initially experimental and later on
complemented by numerical studies. The previous review by Thew (2000) focused mainly
on experimental achievements and remains incomplete, with reference to scant work on
numerical simulations.
The present article provides a literature review of the main findings resulting from
experimental, modeling, and numerical simulation research studies that have taken place
up until 2008 with specific emphasis on chronological development of de-oiling hydro-
cyclone separation technology and issues of cognizance for future research.
Downloaded by [University of Otago] at 07:24 06 October 2014

Evolution of the De-oiling Hydrocyclone Design

Requirements and Flow Features


In the basic design of the de-oiling hydrocyclone (from now on referred to only as a
hydrocyclone for short), the phases to separate enter into the cyclone via a tangential
inlet and by the effect of the wall curvature of the cylindrical body a strongly swirling
stream is developed, causing the lighter phase to migrate toward the centerline and the
heavier one toward the wall (Figure 1). The oil concentration in the inlet stream is usually
between 2,000 to 5,000 ppm. The length of the cylindrical part has to be chosen carefully
because an excessive length leads to weakening the swirling flow due to the wall frictional
effect. Peng et al. (2001) mentioned this in their study of the relationship between the
swirl intensity and the length of a cylindrical cyclone. To avoid the swirl weakening,
the cyclone usually has a conical lower part to maintain the swirl intensity and hence a
constant or stronger centrifugal separation. A swirling reverse flow takes place in the core
region, characterized by a very low pressure, and leaves the cyclone body via a circular
hole (1–4 mm) placed in the center of the upper surface or end wall of the cyclone. The
heavier phase (water rich) is transported by the downward stream, whereas the lighter
(oil rich) one migrates with the upward stream. The de-oiling hydrocyclone works ideally
with a stream of dispersed oil in water and most of the time is unable to handle tight oil
emulsions.
Colman et al. (1984) summarized the requirements for the hydrocyclone; i.e., gener-
ating a high swirl with low shear effect, maintaining the swirl flow for a sufficient time,
minimizing the oil-enriched outlet flow, minimizing the pressure drop, and maintaining
reasonable operating costs. Hydrocyclones have many advantages compared with other
separation devices; namely, short residence time (of the order of 1–2 s), high centrifugal
forces generated (102 –103 times the gravity acceleration), small separated droplet size
(equal or greater than 10 m), compactness (10% of older plant weight and footprint),
low manufacturing and maintenance costs, insensitivity to orientation, a relatively modest
and seldom need for chemicals, and no subsequent processing except when the standards
limitations are not achieved at the underflow outlet of the hydrocyclone (Thew, 2000;
Hashmi, 2005). Unlike solid–liquid hydrocyclones where the dispersed phase is heavier
and is collected at the underflow outlet, the dispersed phase (oil droplets) is lighter
and migrates via the upstream flow so that the product of separation is collected at the
overflow outlet, thus necessitating a special design (Thew, 1986).
740 N. Kharoua et al.
Downloaded by [University of Otago] at 07:24 06 October 2014

Figure 1. Conventional de-oiling hydrocyclone: features and flow topology.

Evolution
The theory and design principle studies of liquid–liquid hydrocyclones began in the
1950s and 1960s (e.g., Bradley, 1965; Rietema, 1961), but the pioneer work was that of
the Southampton University group in the U.K. (Young et al., 1994; Thew, 2000), which
extended over almost four decades.
In the 1960s, a group of researchers from Southampton University, directed by M.
Thew, began a research program to design a hydrocyclone to separate oil, the lighter
phase, from water (Thew, 1986). Their work was inspired by the classical solid–liquid
hydrocyclone designs and motivated by an environmental desire to reduce oil content
in the discharged water from crude oil tankers or produced water in oil fields. The
first and unsuccessful prototype was a cylindrical hydrocyclone design (see Figure 2a)
having tangential inlets and outlets (Colman and Thew, 1980). Experiments employing
polypropylene particles with a mean particle size of 38 m and a density of 900 kg/m3
to simulate oil droplets in hot and cold water were conducted subsequently. The solid
particles experiments have permitted the development of three new designs leading at
the end to one efficient model. Their first efficient and patented hydrocyclone design was
realized in 1978.
Hydrocyclones for De-oiling Applications 741
Downloaded by [University of Otago] at 07:24 06 October 2014

Figure 2. Mutation of Thew’s hydrocyclone design.

The geometry developed by the Southampton University group has evolved over
three phases. Initially, it was a single inlet hydrocyclone with a first cylindrical section
containing an overflow orifice, a second taper section, and a cylindrical tail pipe (Fig-
ure 2b). The geometry was evolved to have a twin tangential inlet with steep and fine
taper sections (Figure 2c). Finally, the tangential inlet was replaced by a larger wrap-
around involute inlet and a shorter cylindrical section, as seen in Figure 2d. Young et al.
(1994) have searched for the optimum hydrocyclone dimensions by studying the effect
of dimensional ratios on single involute hydrocyclone separation efficiency having only
one single taper section. Their proposed design (the Amoco oil–water hydrocyclone),
which represents a modification of the original Colman and Thew design, was shown
to be more efficient than the latter by 20% for droplet size smaller than 20 m. The
modification consisted in adding a vortex finder having virtually the same length as the
first cylindrical section. The cylindrical tailpipe was changed to a very fine taper ending
with a choke.
Although the investigation of the hydrocyclone design had somehow slowed down
after the new millennium, a slightly modified geometry was proposed by Belaidi and
Thew (2003) to enhance the separation of free gas. In this novel geometry, the first
cylindrical section is replaced by a fine taper section (Figure 2e). It was found that the
separation efficiency plateau in the absence of free gas was raised, albeit marginally,
from 98.8%, obtained by the conventional geometry, to 99.45%.
The mini de-oiling hydrocyclone was developed by a research group, formed under a
consortium including industrial partners such as Amoco, Arco, and Chevron, at Michigan
State University directed by C. Petty. Wesson and Petty (1994) have tested an acrylic
Dorr-Oliver 10-mm hydrocyclone with a 132-mm-long steel tailpipe. The mixture was
dispersed kerosene in water with a density difference of 150 kg/m3 and a concentration of
200–300 ppm (measured by laser light scattering). The 10-mm hydrocyclone was capable
of achieving a cut size l50 of 4 m but with the inconvenience of a large split ratio of
50–60%. The mini hydrocyclone was found to be unsuitable for thickening the oily reject
stream (Wesson and Petty, 1994), however. A serial two-stage unit was adopted, where the
742 N. Kharoua et al.

overflow stream of 61 parallel mini hydrocyclones was injected in a 67-mm conventional


hydrocyclone that could not achieve a similar cut size but with its low split ratio was
suitable for the thickening duty.
Hashmi et al. (2004) have presented the performance of the CANMET hydrocyclone
unit patented and introduced for testing in 1999 (Hashmi and Hamza, 2005); see Figure 2f.
Unlike previous designs, the unit had four new features: multiple inlet ports to maintain
a straight oil core, adjustable overflow orifice leading to better controllability, solids
separation attachment, and heating jacket to reduce oil viscosity and hence improve
separation efficiency.
The hydrocyclone diameters are by design of modest sizes. To handle large flow
rates, hydrocyclones have been arranged in parallel bundles (see Meldrum, 1998) where
5 Vortoil units with 35-mm parallel hydrocyclones in each unit processing 5,565 m3 /d of
water. These units are now readily available from industry.
Downloaded by [University of Otago] at 07:24 06 October 2014

Performance of Hydrocyclone and Factors Affecting It


The performance of de-oiling hydrocyclones is measured by the separation efficiency
and its stable operation is influenced by the flow rate through the hydrocyclone Q, the
pressure differential ratio PDR (defined below), and flow split F (defined below). Simple
consideration of Stokes law suggests that the incoming flow characteristics described by
the droplet size distribution, the density differential between the phases and the viscosity,
in addition to the concentration of the secondary phase, and, where applicable, gas content
have a significant effect on performance and yield.
The most important and widely used performance indicator is the separation effi-
ciency (Colman and Thew, 1980; Colman et al., 1980; Meldrum, 1988; Simms et al.,
1992; Gomez et al., 2003). The definition of the separation efficiency, ", varies according
to the hydrocyclone duty (Wolbert et al., 1995; Hashmi et al., 1996; Thew and Smyth,
1998). For the de-oiling duty, the interest is in minimizing the amount of oil present in
the underflow clean water. This parameter can be expressed as follows:
Qo Co
"D (1)
Qi Ci
By applying the mass balance, the separation efficiency expression becomes:
Qi C i Qu C u Qu C u
"D D1 (2)
Qi C i Qi C i
where Qi , Qo , and Qu are the inlet, the overflow, and the underflow flow rates, respectively.
With the assumption that Qo << Qi and hence Qu  Qi , the separation efficiency
expression can be simplified to:
Cu
"D1 (3)
Ci
Oil concentrations and droplet size distribution have been measured in order to
compute the separation efficiency. Previous measurements of concentration were almost
entirely based on isokinetic sampling analyzed by laser light scattering, using the Coulter
counter or infrared-based analyzer (Colman et al., 1980; Simms et al., 1992; Hashmi et al.,
1996; Smyth and Thew, 1996; Gomez et al., 2003). Other methods were also employed.
Hashmi et al. (1996) used liquid nitrogen at 196ı C for a rapid sample freezing to get an
Hydrocyclones for De-oiling Applications 743

image of the sample with confocal laser scanning microscopy in the fluorescence mode.
Belaidi and Thew (2003) have conducted a critical analysis of older methods such as
isokinetic sampling. They mentioned that the photographic and video techniques used in
the 1990s were laborious and presented a wide experimental scatter band. Bennett and
Williams (2004) have employed electrical resistance tomography based on conductivity or
resistivity difference to distinguish between insulating oil and air regions and conducting
water regions inside the hydrocyclone.
Drop size distributions have been measured online using laser interferometry for oil
concentrations of up to 5,000 ppm and ultrasonic interferometry for higher concentrations
(Meyer and Bohnet, 2003).
The typical variation of separation efficiency against flow rate is illustrated in Fig-
ure 3. The higher the flow rate, the higher the swirl intensity, which causes the separation
efficiency to increase rapidly. On the other hand, the shear effect (droplet breakup) also
increases and stops the separation efficiency increase (plateau in Figure 3) at a certain
Downloaded by [University of Otago] at 07:24 06 October 2014

flow rate value. Beyond the flow rate value giving the maximum efficiency, the shear
effect is so important that the separation efficiency drops off sharply. Typical results
describing the relationship between the flow rate and the separation efficiency, displayed
in Figure 3, were carried out by Meldrum (1988) for tests at Murchison and Hutton fields
in the North Sea using a four-parallel-hydrocyclones unit (35-mm-diameter conventional
hydrocyclone) and Bennett and Williams (2004) in a single 70-mm-diameter Mozley’s
design hydrocyclone. Typical values of the turn-down ratio achieved so far are in the
range of 2 up to 7 (if high pressures are available).
A combination of the inlet, overflow, and underflow outlet pressures has given a
control parameter called pressure drop ratio (PDR). It is the ratio of the difference
between the inlet and overflow pressures to the difference between the inlet and underflow
pressures.
Pi Po
PDR D (4)
Pi Pu
Usually a minimum underflow pressure is needed as a back-pressure to force the
oil-rich core out from the overflow orifice (Young et al., 1994; Caldentey et al., 2002).
Young et al. (1994) have studied the effects of the underflow pressure and the pressure

Figure 3. Typical separation efficiency vs. flow rate curve with lines representing main trend.
744 N. Kharoua et al.

drop between the inlet and the underflow outlet on the separation efficiency for Bumpass
crude (0.85 g/cm3 ) and South China Sea (0.95 g/cm3 ) in a 35-mm Colman and Thew
hydrocyclone. The flow rate was found to depend significantly on the pressure drop
in agreement with the theoretical relation giving a parabolic trend. In addition, with a
minimum of underflow back-pressure between 4.14 and 5.17 bar, the minimum overflow
volume percentage flow rate required to reach maximum efficiency is around 1.5%. In
general, the effects of inlet, underflow, and overflow pressures are controlled by adjusting
the PDR value. Typical suggested values of PDR are in the range of 2–3 (Thew and
Smyth, 1998; Belaidi and Thew, 2003).
Although the overflow rate represents only a few percent of the inlet flow rate,
it has the important role of maintaining the stability of the core stream. The relevant
parameter is the flow split ratio (F) defined as the ratio of the overflow flow rate to the
inlet flow rate. Below a minimum value, some oil begins to be lost in the downstream
flow and the separation efficiency decreases. On the other hand, increasing F beyond the
Downloaded by [University of Otago] at 07:24 06 October 2014

minimum value implies that more water is present in the overflow (Colman and Thew,
1980; Colman et al., 1980; Colman et al., 1984; Meldrum, 1988; Young et al., 1994).
Typically for a good design F varies in the range of 2–3%. Meldrum (1988) mentioned
that a value of F equal to 1% for an efficiency plateau of 85% was sufficient during the
Hutton and Murchison field tests cited previously.
Because the separation in hydrocyclones is based on the effects of centrifugal and
centripetal forces, the smaller (lighter) oil droplets are less sensitive to the continuous
phase flow behavior and hence to the separation action. Young et al. (1994) presented
results of underflow efficiency, defined in Eq. (3), versus overflow volume percent for
a 35-mm-diameter hydrocyclone using Bumpass and South China sea crude varying the
split ratio between 0.5 and 2.5% for several oil droplet diameters (d50 ) ranging from 15
to 55 m. The separation efficiency was improved from a plateau of 50% for a fine
average droplet size or around 15 m to 98% for a larger average droplet size of 55 m.
Thus, to increase the droplet size in order to improve separation efficiency, a treatment
or preconditioning of the feed was applied in some de-oiling installations. Sinker et al.
(1999) have presented the performance enhancement coalescence technology (PEFC-F)
installation where a precoalescer has been placed upstream of the hydrocyclone. The tests
were conducted at a temperature of 30ı C, a droplet size range of 8 m d.50/, 18 m d.90/,
and an inlet oil concentration range of 200 to 550 ppm. Comparative results obtained for
hydrocyclone operating with the coalescer offline and online with an increasing pressure
drop between the inlet and the overflow outlet from 1 to 8 bar showed an improvement of
the separation efficiency from 43 to 63% for the offline coalescer tests, whereas for the on-
line coalescer case, the separation efficiency was stabilized at a value around 91%. Hashmi
et al. (2004) have obtained improved overflow separation efficiency by conditioning slop
oil with mean droplet size of 4 m. Thus, hot water at 87ı C was added to the crude
and by applying a slow-speed mixing led to a larger mean droplet size of 10 m. Tests
with and without conditioning have been conducted in a two-stage unit. The separation
efficiency at the second stage jumped from 20 to 90% due to the droplets growth.
The separation efficiency is slightly improved when the inlet oil concentration in-
creases. Young et al. (1994) have concluded that all the additional oil could be removed
if there is adequate split ratio to remove the required amount of oil. They found that
the separation efficiency remained constant (around 90%) over a wide range of inlet
oil concentrations between 0 and 4,000 ppm. Colman et al. (1980) have carried out
experiments for concentrations starting from 100 ppm and reaching 30,000 ppm. The
tests were conducted using a 58-mm-diameter hydrocyclone with two different mean
Hydrocyclones for De-oiling Applications 745

droplet diameters in the feed (41 and 85 m) and a 30-mm-diameter hydrocyclone with
a mean droplet diameter of 41 m in the feed. For each of the cases cited, the separation
efficiency remained constant over all the ranges of concentrations tested.
By heating the mixture, the oil viscosity is reduced and the droplets become more
sensitive to the continuous phase flow effects. Hashmi et al. (2004) have tested the
effects of heating on Canadian heavy oil crude with the CANMET hydrocyclone unit.
The separation efficiency was improved from 55.5% for 20ı C to an average of 98% for
40ı C. Between 40ı C and 60ı C there was no significant improvement.
Investigations on the presence of both free and dissolved gas have demonstrated
that the presence of high fractions of free or dissolved gas could be disruptive. Smyth
and Thew (1996) have used an oil/water mixture saturated by 68% of dissolved gas. To
control the effects of the presence of dissolved gas, they recommended an overflow orifice
diameter up to 3 mm and explained what the optimum operational conditions should be;
i.e., highest possible pressures and lowest suitable pressure drops and flow splits. Belaidi
Downloaded by [University of Otago] at 07:24 06 October 2014

and Thew (2003) found that the free gas became detrimental when its concentration in the
inlet stream exceeded 40% and increased exponentially when comparing the separation
efficiency to a case without free gas. Indeed, the reverse flow is dominated by a gas
core surrounded by an annulus of liquid. Thus, an excessive amount of gas caused the
reduction of the area occupied by the liquid, leading to choking. They proposed a modified
geometry where the first cylindrical section in the conventional geometry was replaced
by a fine taper section to cause more rapid separation of free gas in this section.

Geometrical Features of Hydrocyclones


As discussed previously, the present-day designs of de-oiling hydrocyclones have seen
considerable mutations in size and geometrical features over the past 25 years. These
geometrical features, depicted in Figure 4, including the inlet, the cylindrical section, the
taper section, tailpipe, and vortex finder are now reviewed in turn. Table 1 summarizes

Table 1
Geometrical dimensions of models
Dtap Do Di Dt Ls Lt L ˛ı ˇı ı

0.5 0.07 0.175 (twin 0.25 1 10 22.5 0 10 0.75 Colman and


inlet) Thew (1983)
0.75 0.07 (Ds /12)  0.5s 0.25 0.5 10 21 0 10 0.75 Hargreaves and
Single Silvester (1990)
rectangular
inlet
— 0.039 0.25 0.33 0–2 9 — 0 6 — Young et al.
(1994)
— 0.24 0.24 0.22 1 13.5 21 0 6 — Wesson and Petty
(1994), mini
hydrocyclone
Ds  10 mm
0.75 — Larger than 0.375 0.375 10 26–33 0 10 0.75 Thew (2000),
the twin- single involute
inlet inlet
design
0.475 0.2–0.6 0.15 0.24 0.36 4.7 13 <10 10 1 Belaidi and Thew
(2003)
746 N. Kharoua et al.
Downloaded by [University of Otago] at 07:24 06 October 2014

Figure 4. Geometrical features of the de-oiling hydrocyclone.

the evolution of the characteristic dimensions normalized by the reference diameter of


the Southampton conventional hydrocyclone design and its derivatives.

Inlet
The size and the shape of the inlet have a considerable effect on the flow field behavior.
Young et al. (1994) have found that a ratio of the inlet diameter to the reference
diameter (diameter of the two taper sections junction) equal to 0.25 was the optimum
and larger ratios could give better separation efficiency but needed higher flow rates.
The Southampton group used a ratio of 0.35 for their twin-inlet hydrocyclone. Another
design of the inlet part is the rectangular tangential inlet introduced through a gentle
volute (Hargreaves and Silvester, 1990; Thew, 2000). The swirl number defined as a
Hydrocyclones for De-oiling Applications 747

function of a characteristic inlet area and the hydrocyclone reference diameter was found
to be in the range 8–10. It reflects the effects of the inlet diameter on the swirl intensity.
This part of the hydrocyclone is known to be a high-turbulence region and an investigation
of its shape can permit a further improvement of the separation efficiency by reducing
the shear effects (see Small et al., 1996).

Cylindrical Section
A cylindrical section is necessary to avoid a high shear region downstream of the entry
and to reduce the head loss. Colman et al. (1984) explained that to generate a similar
high swirl leading to a swirl number in the range of 8–10 without using a cylindrical
section, a smaller hydrocyclone with small inlets was required. This was confirmed by
experimental tests. Although the presence of a cylindrical section is indispensable, it must
be as short as possible (Young et al., 1994). Indeed, the involute hydrocyclone design of
Downloaded by [University of Otago] at 07:24 06 October 2014

the Southampton group has a shorter cylindrical part with a length equal to the reference
diameter instead of two times the reference diameter used in the twin inlet design.

Taper Sections
Two options are available; single-cone or bi-cone hydrocyclones. For the single-cone
design a value of 6ı was recommended by Young et al. (1994) as an optimum cone
angle. The Southampton group bi-cone hydrocyclone had a first short, steep taper section
with an angle of 20ı , which had the role of accelerating the swirling stream, and a second,
longer, fine taper section with an angle of 1.5ı to increase the residence time without
damping the swirling flow and hence increasing the chance of separating more droplets.

Tailpipe
A prolonged tailpipe connected to the last taper section allows the separation of finer
droplets that escape the swirl effect in the upper regions. However, a very long tailpipe
(more than 40 or 50 times the reference diameter) impairs the hydrocyclone advantage
of compactness. The Southampton group design tailpipe was 20 times the reference
diameter. Young et al. (1994) have replaced the cylindrical tailpipe with a very slightly
tapered section (cone angle of 1 to 15 min!) to compensate the frictional loss by a slight
acceleration.

Vortex Finder
Because the oil droplets are collected at the overflow orifice after migrating toward
the central region, it is not expected that they could be lost in short circuits; thus, a
projecting vortex finder is redundant in the case of the de-oiling hydrocyclone. However,
in the improved hydrocyclone of Young et al. (1994), a vortex finder is projected until the
junction between the cyclindrical section and the first taper section. Colman and Thew
(1980) have designed a coaxial outlet hydrocyclone based on the fact that the upstream
flow was constituted by an oil core surrounded by water. When adding a coaxial outlet,
the amount of water present in the overflow stream was collected in the annular space
reducing the water amount from the oil-rich stream. The overflow orifice size determines
the operating flow splits and might be a remedy to the overflow choking by the presence
of gas. In fact, Belaidi et al. (2003) used several overflow orifice sizes (0.2–0.6 times the
748 N. Kharoua et al.

reference diameter) to eliminate the effect of high amounts of free gas on the hydrocyclone
performance.
The Southampton University design remains the typical hydrocyclone largely used
in industry. Although recent studies are focusing on the de-oiling units design rather
than the hydrocyclone design, the modifications proposed in the 2000s (e.g., Belaidi
and Thew, 2003; Hashmi and Hamza, 2005) indicate that an eventual improvement
of the hydrocyclone geometry is still possible. The actual trend is to minimize the
hydrocyclone diameter because smaller diameters are more suitable for generating higher-
speed swirling flows. Thew (2000) reported that the earliest hydrocyclones had, in general,
relatively large diameters (60–140 mm) compared with those manufactured recently (30–
60 mm). The use of smaller hydrocyclones (mini-hydrocyclones) is avoided because of
the disadvantages cited in previously.

FLUID Mechanics of De-oiling Hydrocyclones


Downloaded by [University of Otago] at 07:24 06 October 2014

The flow field inside a de-oiling hydrocyclone is quite complex through the combination
of time-dependent, turbulent, three-dimensional, high-intensity swirl and multiphase flow
with interaction of two or three phases, including droplet breakup and coalescence.
Knowledge of the flow behavior inside the de-oiling hydrocyclone is fundamental for
understanding, predicting, and improving its performance. Previous works on the fluid
mechanics included limited experimental and theoretical studies and computational fluid
dynamics (CFD) studies.
Despite the fact that understanding the flow field development inside a de-oiling
hydrocyclone is necessary for design improvements and optimization of performance,
only few detailed investigations have been conducted that treated the flow dynamics
inside de-oiling hydrocyclones (Colman et al., 1984; Lu et al., 1997; Lu and Zhou,
2003; Bai et al., 2008). This was made possible with nonintrusive optical laser-based
instrumentation. Colman et al. (1984), using a nominal size of 116 mm, have published
radial profiles of the axial velocity component at eight axial positions for a twin-inlet
hydrocyclone. The results identified two regions separated by a zero axial velocity value
boundary. In the outside region near the wall, a downward flow was accelerated in the
reduced hydrocyclone axial cross sections. In the inner core, an upward flow accelerated
when getting closer to the overflow orifice. At the end of the tail pipe section no negative
axial velocities were seen, which means that no reversal flow occurred at that position.
The location of the beginning point of the reversal stream and its length, also known as
the natural length of the hydrocyclone, can be extracted if the deepest axial position of
nonnegative axial velocity is accurately interpolated and is found to be virtually equal
to the whole of the hydrocyclone length. However, unlike the conventional de-oiling
hydrocyclone, Bai et al. (2008) found a length of 200 mm out of a hydrocyclone total
length of 632.8 mm with no separation occurring in the tailpipe. The profiles of the
tangential velocity remain similar inside the cylindrical part and the conical part, where
the maximum values are measured and these decrease along the axial direction in the
tailpipe due to the frictional effect on the swirl decay. The maximum tangential velocities
separating the two regions of quasi-forced vortex and quasi-free vortex, observed in
different axial positions, increased with the hydrocyclone height. Their radial positions
were changing in a thin core region (no more than one sixth of the radius at each axial po-
sition). The maximum values of axial velocity fluctuations were found to be in the conical
portion close to the wall due to the high mean velocity gradients existing in this region
and represent approximately 4.8% of the average inlet velocity. The maximum tangential
Hydrocyclones for De-oiling Applications 749

velocity fluctuations were observed near the wall and the centerline because, unlike for the
axial velocity fluctuations, high mean velocity gradients are generated in the centerline
region inside the quasi-forced vortex also (Colman and Thew, 1980; Caldentey et al.,
2002; Bai et al., 2008). The radial mean velocity component, usually negligible compared
to the other components, is usually difficult to measure. In theoretical studies (see Wolbert
et al., 1995; Caldentey et al., 2002) where an expression for its profile is derived from the
continuity equation and the wall condition suggested by Kelsall (1952). The signal quality
of a laser Doppler anemometer near the central axis can deteriorate because of lower
seeding density in this region. And although point measurements using laser Doppler
anemometry were reported previously, particle image velocimetry (PIV), a nonintrusive
technique that allows quantification of global aspects of the flow, has not been used so far.
Different theoretical models have been developed to predict the separation efficiency.
Martins et al. (1996) presented a model where a global efficiency was defined as a
function of the split ratio and the diameters having zero value migration probability
Downloaded by [University of Otago] at 07:24 06 October 2014

and reduced migration probability, respectively. An interesting model is that of Wolbert


et al. (1995) based on trajectory analysis. The authors presented a critical discussion
of older separation efficiency models for solid–liquid hydrocyclones; namely, the radial
equilibrium locus theory and the residence time theory, both unfavorable for the liquid–
liquid case. The idea of trajectory analysis was also used by Moraes et al. (1996) and
Caldentey et al. (2002). These approaches based on oversimplifying assumptions lead to
unrealistic and unreliable results.
In contrast to solid–liquid hydrocyclones, simulations of oil–water hydrocyclones are
somewhat lagging behind. CFD can offer the advantage of providing short turnaround
times during design and provides a cost-effective analysis tool. In CFD, the first basic
and classical approach is to solve the Reynolds-averaged Navier-Stokes equations in
combination with a turbulence model. The k-" model based on an isotropic turbulent
viscosity and equilibrium assumptions is now generally recognized as unsuitable for
modeling the highly anisotropic turbulence in hydrocyclones. The turbulence anisotropy
is better predicted by a stress transport model, either in its algebraic or differential
form (e.g., RSM LRR of Launder et al. (1975), RSM SSG of Speziale et al. (1991)),
which solves a transport equation for each individual Reynolds stress. Better results were
obtained with quadratic pressure strain correlation (RSM SSG).
At a second level, the large eddy simulation (LES) model known to be more accurate
compared with the k-" or the RSM models (Delgadillo and Rajamani, 2005) has become
the reference turbulence model because of the remarkable advances in computational
hardware resources, allowing more complex-turbulence model simulations in very fine
meshes and in an acceptable CPU time.
At the other extreme of complexity, direct numerical simulation (DNS) offers the
most fundamental approach because it attempts to resolve all of the time and length scales
of the flows and therefore puts a formidable requirement on computational resources for
high Reynolds number flows of industrial relevance. Its use has therefore been limited
to simple low Reynolds number flows and would therefore remain out of the scope for
real flows.
In addition to the turbulence model, it is also now well acknowledged that a second-
order numerical discretization scheme is needed to attenuate numerical discretization
errors and diffusion and particular attention has to be paid to the mesh topology and grid
distribution, especially near solid walls and when a wall function model is used or for
high-gradients regions. A successful CFD computation will also require a correct and
physically realistic set of boundary conditions at the inlet and outlets.
750 N. Kharoua et al.

To completely capture the multiphase nature of flows in hydrocyclones, a multiphase


flow model is needed. There are two fundamental approaches available, the Lagrangian
approach based on tracking particles trajectories and the Eulerian approach, which in-
cludes as subsets the full Eulerian multiphase model, the mixture model, and the volume
of fluid model. A comparative study between the Lagrangian approach and the Eulerian
approach has been conducted by Durst et al. (1984). The Lagrangian approach solves an
integrated equation of the force balance on particles, bubbles, or droplets subsequent to
the single-phase solution of the flow field. It has the advantage to predict more details
of the dispersed phase such as the trajectories and the residence times. It performs better
for flows with higher particle velocity gradients and is less sensitive to numerical errors.
Furthermore, the Lagrangian approach is less time consuming and more appropriate for
polydispersed secondary phases. The Eulerian approach considers the phases as interpen-
etrating continua and solves the continuity and momentum equations for each phase. A
simplified form of the Eulerian model is the mixture model, which solves the continuity
Downloaded by [University of Otago] at 07:24 06 October 2014

and momentum equations for the mixture and a transport equation for the volume fraction
of the secondary phases (Manninen et al., 1996). An algebraic slip-velocity formulation
is adopted to take into account the relative velocity of the secondary phases. The Eulerian
approach predicts better flows with suspended dispersed phases (large residence time)
and is the alternative of the Lagrangian approach when high loadings cause solution in-
stabilities. In addition, the Eulerian approach permits simulating the effects of turbulence
on the dispersed phase. Generally, a threshold of 10% for the secondary phase loading by
volume necessitates passing from the Lagrangian to the Eulerian approach. The particle
loading and the Stokes numbers are introduced as useful parameters when choosing a
multiphase model for intermediate loadings. For high loading, the use of the Eulerian
model is indispensable (Durst et al., 1984; Fluent version 6.3, 2006 documentation).
The important phenomena of droplets breakup and coalescence were omitted in all of
the previous studies due to the complicated flow structure in a liquid–liquid hydrocyclone.
They certainly deserve research effort in the future.
The work of Hargreaves and Silvester (1990) deserves mention because these authors
were the first to investigate a de-oiling hydrocyclone using CFD. The geometry employed
was that of a single inlet Colman and Thew hydrocyclone. They used the algebraic stress
model with several oversimplifying assumptions. A Lagrangian droplet tracking model
using the particle motion simulation basis was provided by Boysan et al. (1982) with
the following assumptions: no droplet–droplet interaction, no influence of droplet slip on
the bulk flow field, no coalescence or disintegration of droplets, a Gaussian distribution
of fluctuating velocity components, no added mass effects for drag-accelerated particles,
and no mass transfer to or from a droplet. Because the effect of the continuous phase on
the droplet could be more important than that on solid particles, the authors integrated a
weighting factor to the fluctuating velocity components in the set of momentum equations
of droplet motion. The numerical results exhibited an overestimation of the axial velocity
and an overestimation of the tangential velocity, which remained constant in different axial
positions contrary to the experimental case. Overall the simulation predicted mitigated
results. The migration probability curves showed an overestimation compared with the
experimental results of Nezhati and Thew (1987).
Grady et al. (2003) studied the flow field and the performance of a mini-hydrocyclone
(10 mm diameter) similar in shape to the involute inlet of Colman and Thew model with
only one fine taper section (angle 1.8ı). The turbulence model used was the Reynolds
stress model of Launder et al. (1975). The main results obtained were an asymmetry
caused by the single inlet and a reverse flow occurring in the forced vortex region
Hydrocyclones for De-oiling Applications 751

and extending almost to 80% of the hydrocyclone length. They used the algebraic
mixture model implemented in the Fluent software with the simplifying hypotheses that
the droplets were immiscible, noncoalescent, and nondeliquescent. Their results were
characterized by a good prediction of the separation efficiency for droplet sizes larger
than 20 m. The simulation did not take into account the breakup phenomenon leading
to an additional population of fine droplets in the underflow outlet and represented by
negative separation efficiencies. This caused the simulation results to become unreliable
in predicting the separation efficiency for small droplets compared with experiments.
In a different attempt and strategy, Petty and Parks (2004) solved the single-phase
Navier-Stokes equations by DNS, for a laminar flow, in a cylindrical mini-hydrocyclone
(5 mm diameter) divided into 135,000 finite volumes. The classical results for the flow
field in a conventional hydrocyclone, a Rankine vortex structure for the tangential velocity
and a reversal flow in the axial direction reaching the bottom of the mini-hydrocyclone
were obtained but without comparison with experimental results.
Downloaded by [University of Otago] at 07:24 06 October 2014

The mixture multiphase model was utilized by Paladino et al. (2005). This model
can account for populations of several classes of droplet sizes, six in this case, at a
reasonable computational cost. Contours of the oil volume fraction distribution were
presented for each class and for the total population. The results show that the individual
group separation efficiencies were very poor except for the group having the mean droplet
size 165.49 m. The global efficiency of the whole population was about 35%. This
numerical result agrees qualitatively with the experimental results (e.g., Young et al.,
1994; Sinker et al., 1999; Hashmi et al., 2004), showing an improvement of the separation
efficiency with the increase in the droplet size.
Huang (2005) combined the Reynolds stress turbulence model with the more accurate
and complicated Eulerian-Eulerian model to simulate the multiphase flow inside a 30-mm-
diameter Thew hydrocyclone with twin square inlets. The oil concentration was over 10%
by volume and the split ratio was 10%. The author tried to illustrate the development of
the separation mechanism with time. The use of LES is still in its beginning stage. Slack
et al. (2006) presented, briefly, results of the flow field simulation inside a CANMET
de-oiling hydrocyclone, but no published research work is available that is backed by
experimental results. The LES model remains the most promising model used recently
for both flow field and classification efficiency predictions.
Future studies will be challenged by minimizing the simplifying assumptions to take
account of the droplet reaction to the effect of the highly swirling turbulent flow inside
the de-oiling hydrocyclone, which causes the oil-dispersed phase at the inlet to become
a continuous phase at the core region due to coalescence and a finer dispersion at the
underflow outlet due among other things to breakup. A similar behavior concerning air–
liquid multiphase flow has been judged susceptible to be simulated using the mixture
model (Brennan, 2006; Kharoua et al., 2009). More detailed experimental studies of
the different phenomena occurring inside the hydrocyclone, such as core formation,
droplet behavior, and turbulence effects on these two phenomena will probably enrich the
internal flow behavior knowledge and hence reduce the need of simplifying assumptions.
In addition, the use of benchmark detailed and complete experimental results should
allow the use of more realistic boundary conditions at both the inlets and outlets of
the hydrocyclone. The use of more robust and accurate models for both turbulence and
multiphase flow prediction is another challenge due to the complexity of the hydrocyclone
geometry and the flow field requiring a very fine mesh (few million cells). This is being
overcome by the remarkable advances in computational hardware.
752 N. Kharoua et al.

Conclusions
The hydrocyclone technique for de-oiling produced water from oil fields or slops has
become one of the most efficient and proven oil/water separation technology due to its
simplicity, compactness, robustness, and low manufacturing and maintenance costs. Their
performance is still subject to drop size and feed concentration limitations.
Although hydrocyclones have been well studied experimentally, some phenomena
are still poorly understood and need more attention; namely, the effects of turbulence
on droplets motion, breakup and coalescence and their spatial distribution, central core
shapes, dimensions, behavior, and the effect of the presence of gas (dissolved or free) or
solids on the flow field.
It is noteworthy to mention that most of the de-oiling hydrocyclone investigations
were experimental in nature and very few works have explored the complex flow field
inside the hydrocyclone by both experiments and CFD. Although the experimental studies
have permitted to some extent characterizing the hydrocyclone performance, the benefits
Downloaded by [University of Otago] at 07:24 06 October 2014

that can be obtained using CFD are numerous: low turnaround time, low cost, and ability
to explore several geometrical variants with low cost and effort. Furthermore, the actual
powerful means of computation allow the gradual introduction of more complex and
accurate models, thus providing more reliable results.
Several of the limited studies that utilized CFD applied several simplification hy-
potheses such as axisymmetry, k-" turbulence model, the omission of droplet breakup or
coalescence effects, admittedly, to reduce the time and the cost of computation, which
led, however, to unreliable results. There is now agreement on the necessity to use more
accurate models for both turbulence and multiphase behavior. For turbulence, at least
a full Reynolds stress or an LES model is required. The multiphase or discrete phase
model used should take into account the fact that two different multiphase flow regimes
are present at the same time; dispersed oil droplets at the hydrocyclone inlet a quasi-
continuous phase due to coalescence (stratified flow) and that droplets break up may exist
depending on the local flow conditions. There is hence additional scope for successful
application of CFD modeling techniques to better understand the flow behavior inside
the hydrocyclone.

References
ANSYS, Inc. (2006). Fluent 6.3 Documentation, User’s Guide. Canonsburg, PA: Fluent Inc.
Bai, Z. S., Wang, H. L., and Tu, S. T. (2008). Experimental study of flow patterns in de-oiling
hydrocyclone. Miner. Eng. 22:319–323.
Belaidi, A., and Thew, M. T. (2003). The effect of oil and gas content on the controllability and
separation in a de-oiling hydrocyclone. Trans. Inst. Chem. Eng. 81:305–314.
Bennett, M. A., and Williams, R. A. (2004). Monitoring the operation of an oil/water separator
using impedance tomography. Miner. Eng. 17:605–614.
Bhaskar, K. U., Murthy, Y. R., Raju, M. R., Tiwari, S., Srivastava, J. K., and Ramakrishnan, N.
(2007). CFD simulation and experimental validation studies on hydrocyclone. Miner. Eng.
20:60–71.
Boysan, F., Ayers, W. H., and Swithenbank, J. (1982). A fundamental mathematical modelling
approach to cyclone. Trans. Inst. Chem. Eng. 60:222–230.
Bradley, D. (1965). The Hydrocyclone. London: Pergamon Press.
Brennan, M. (2006). CFD Simulation of hydrocyclones with an air core: Comparison between
large eddy simulations and a second moment closure. Chem. Eng. Res. Des. 84:495–505.
Hydrocyclones for De-oiling Applications 753

Caldentey, J., Gomez, C., Wang, S. B., Gomez, L., Mohan, R., and Shoham, O. (2002). Oil/water
separation in liquid/liquid hydrocyclones (LLHC): Part 2—Mechanistic modeling. SPE J.
7:362–372.
Colman, D. A., and Thew, M. T. (1980). Hydrocyclone to give a highly concentrated sample of
a lighter dispersed phase. Proceedings of the International Conference on Hydrocyclones,
Cambridge, UK, October 1–3, pp. 209–223.
Colman, D. A., and Thew, M. T. (1983). Correlation of separation results from light dispersion
hydrocyclones. Chem. Eng. Res. Des. 61:233–240.
Colman, D. A., Thew, M. T., and Corney, D. R. (1980). Hydrocyclones for oil/water separation.
International Conference on Hydrocyclones, Cambridge, UK, October 1–3, pp. 143–165.
Colman, D. A., Thew, M. T., and Lloyd, D. D. (1984). The concept of hydrocyclones for separating
light dispersions and a comparison of field data with laboratory work. Paper no. F2, Proceed-
ings of the 2nd International Conference on Hydrocyclones, Bath, England, September, 19–21,
pp. 217–232.
Delgadillo, J. A., and Rajamani, R. K. (2005). A comparative study of three turbulence-closure
models for the hydrocyclone problem. Int. J. Miner. Process. 77:217–230.
Downloaded by [University of Otago] at 07:24 06 October 2014

Durst, F., Milojevic, D., and Schonung, B. (1984). Eulerian and Lagrangian predictions of partic-
ulate two-phase flows: A numerical study. Appl. Math. Model. 8:101–112.
Gomez, C., Caldentey, J., Wang, S. B., Gomez, L., Mohan, R., and Shoham, O. (2003). Oil/water
separation in liquid/liquid hydrocyclones (LLHC): Part 1—Experimental investigation. SPE J.
7:353–361.
Grady, S. A., Wesson, G. D., Abdullah, M., and Kalu, E. E. (2003). Prediction of 10-mm hydro-
cyclone separation efficiency using computational fluid dynamics. Filtrat. Separ. 40:41–46.
Hargreaves, J., and Silvester, R. (1990). Computational fluid dynamics applied to the analysis of
de-oiling hydrocyclone performance. Trans. Inst. Chem. Eng. 68:365–383.
Hashmi, K. A. (2005). CD of the workshop on separation methods: Kuwait oil sector: December.
19–21.
Hashmi, K. A., Friesen, W. I., Bohun, D. A., and Thew, M. T. (1996). Application of hydrocy-
clones for treating produced fluids in heavy oil recovery. In: Hydrocyclones ’96, Claxton, D.,
Svarovsky, L., and Thew, M. (Eds.), London: Mechanical Engineering Publications Limited,
pp. 357–368.
Hashmi, K. A., and Hamza, H. A. (2005). Integration of CANMET hydrocyclone in conventional
heavy oil treatment facility. Journal of Canadian Petroleum Technology, 44:12–15.
Hashmi, K. A., Hamza, H. A., and Wilson, J. C. (2004). CANMET hydrocyclone: An emerging
alternative for the treatment of oily waste streams. Miner. Eng. 17:643–649.
Huang, S. (2005). Numerical simulation of oil–water hydrocyclone using Reynolds-stress model
for Eulerian multiphase flows. Can. J. Chem. Eng. 83:829–834.
Kelsall, D. F. (1952). A study of the motion of solid particles in or in terms of the volumetric flow
rate as a hydraulic cyclone. Trans. Inst. Chem. Eng. 30:87–104.
Kharoua, N., Khezzar, L., and Nemouchi, Z. (2009). Simulation of the effects of free gas on the
flow behavior inside a liquid–liquid hydrocyclone. Proceedings of the Fourth International
Conference on Thermal Engineering: Theory and Applications, Paper No. 64, Abu Dhabi,
UAE, January 12–14.
Launder, B. E., Reece, G. J., and Rodi, W. (1975). Progress in the development of a Reynolds-stress
turbulence closure. J. Fluid Mech. 68:537–566.
Lu, Y. J., Shen, X., and Zhou, L. X. (1997). LDV Diagnostics of the flow field in a hydrocyclone.
Acta Mechanical Sinica 29:395–405.
Lu, Y., and Zhou, L. (2003). Numerical simulation of fluid flow and oil-water separation in
hydrocyclone. Chinese J. Chem. Eng. 11:97–101.
Manninen, M., Taivassalo, V., and Kallio, S. (1996). On the Mixture Model for Multiphase Flow.
Espoo, Finland: Valtion Teknillinen Tutkimuskeskus.
Martins, R. M. L., Nunes Dias, C. A., and Feres, A. N. (1996). A theoretical–experimental method
for analysis of hydrocyclones for treating oily waters. In: Hydrocyclones ’96, Claxton, D.,
Svarosky, L., and Thew, M. (Eds.), London: Mechanical Engineering Publications Limited,
pp. 333–344.
754 N. Kharoua et al.

Meldrum, N. (1988). Hydrocyclones: A solution to produced-water treatment. SPE Prod. Eng.,


November: 669–676.
Meyer, M., and Bohnet, M. (2003). Influence of entrance droplet size distribution and feed con-
centration on separation of immiscible liquids using hydrocyclones. Chem. Eng. & Tech.
26:660–665.
Moraes, C. A. C., Hackenberg, C. M., Russo, C., and Medronho, R. A. (1996). Theoretical analysis
of oily water hydrocyclones. In: Hydrocyclones ’96, Claxton, D., Svarosky, L., and Thew, M.
(Eds.), London: Mechanical Engineering Publications Limited, pp. 383–398.
Nezhati, K., and Thew, M. T. (1987). Aspects of the performance and scaling of hydrocyclones
for use with light dispersions. Paper no. G1, 3rd International Conference on Hydrocyclones,
Oxford, England, September 30, pp. 167–180.
Paladino, E. E., Nunes, G. C., and Schwenk, L. (2005). CFD analysis of the transient flow in a
low-oil concentration hydrocyclone. Conference Proceedings of the AIChE Annual Meeting,
Cincinnati, OH, October 30–November 4, pp. 646–657.
Peng, W., Boot, P., Udding, A., Hoffman, A. C., Dries, H. W. A., Ekker, A., and Kater, J.
(2001). Determining the best modelling assumptions for cyclones and swirl tubes by CFD and
Downloaded by [University of Otago] at 07:24 06 October 2014

LDA. Paper No. 113 International Congress for Particle Technology, Nuremberg, Germany,
March 27–29, pp. 1–8.
Petty, C. A., and Parks, S. M. (2004). Flow structures within miniature hydrocyclones. Miner. Eng.
17:615–624.
Rietema, K. (1969). Performance and design of hydrocyclones:I, II, III. Chem. Eng. Sci. 15:298–325.
Simms, K. M., Zaidi, S. A., Hashmi, K. A., Thew, M., and Smyth, I. C. (1992). Testing of the
vortoil deoiling hydrocyclone using Canadian offshore crude oil. 4th International Conference
on Hydrocyclones, Southampton, UK, September, pp. 295–308.
Sinker, A. B., Humphris, M., and Wayth, N. (1999). Enhanced de-oiling hydrocyclone performance
without resorting to chemicals. SPE Paper no. 56969, Offshore Europe Conference, Aberdeen,
Scotland, September 7–9, pp. 1–9.
Slack, M., Vedapuri, D., Hashmi, K. A., and Hamza, H. A. (2006). CFD modeling of CANMET
hydrocyclone for the oil industry. 2nd International Symposium on Hydrocyclones, Falmouth
Beach Resort Hotel, Falmouth, UK, June 14–15.
Small, D. M., Fitt, A. D., and Thew, M. T. (1996). The influence of swirl and turbulence anisotropy
on CFD modelling for hydrocyclones. In: Hydrocyclones ’96, Claxton, D., Svarosky, L., and
Thew, M. (Eds.), London: Mechanical Engineering Publications Limited, pp. 49–61.
Smyth, I. C., and Thew, M. T. (1996). A study of the effect of dissolved gas on the operation of
liquid–liquid hydrocyclones. In: Hydrocyclones ’96, Claxton, D., Svarosky, L., and Thew, M.
(Eds.), London: Mechanical Engineering Publications Limited, pp. 357–368.
Speziale, C. G., Sarkar, S., and Gatski, T. B. (1991). Modelling the pressure–strain correlation of
turbulence: An invariant dynamical systems approach. J. Fluid Mech. 227:245–272.
Thew, M. T. (1986). Hydrocyclone redesign for liquid–liquid separation. Chem. Eng. July/August:
17–23.
Thew, M. T. (2000). Cyclones for oil/water separation. In: Encylopedia of Separation Science,
Wilson, I. D. (Ed.), New York: Academic Press, pp. 1480–1490.
Thew, M. T., and Smyth, I. C. (1998). Development and performance of oil–water hydrocyclone
separators—A review. In: Innovation in Physical Separation Technologies, London: The In-
stitution of Mining and Metallurgy, pp. 77–89.
Wesson, G. D., and Petty, C. A. (1994). Process engineering of produced water treatment facility
based on hydrocyclone technology. Proceedings of the International Petroleum Environmental
Conference, Houston, TX, March 2–4, pp. 110–120.
Wolbert, D., Aurelle, Y., and Seureau, J. (1995). Efficiency estimation of liquid–liquid hydrocy-
clones using trajectory analysis. AICHE J. 41:1395–1402.
Young, G. A. B., Taggart, D. L., Hild, D. G., Simms, D. W., and Worrell, J. R. (1993). Hydrocyclone
with Finely Tapered Tail Section. U.S. Patent No. 5,225,082.
Young, G. A. B., Wakley, W. D., Taggart, D. L., Andrews, S. L., and Worrell, J. R. (1994). Oil–water
separation using hydrocyclones—An experimental search for optimum dimensions. J. Petrol.
Sci. Eng. 11:37–50.
Hydrocyclones for De-oiling Applications 755

Nomenclature
Ai total inlet area, mm2
Ci oil concentration at the inlet, ppm or %
Co oil concentration at the overflow, ppm or %
Cu oil concentration at the underflow, ppm or %
Di inlet diameter, mm
Do overflow outlet diameter, mm
Dt tailpipe diameter, mm
Dtap diameter at the junction between the two taper sections, mm
L hydrocyclone length, mm
Ls swirl chamber length, mm
Lt tailpipe length, mm
Qi inlet flow rate, m3 /s
Qo upper outlet flow rate, m3 /s
Downloaded by [University of Otago] at 07:24 06 October 2014

Qu lower outlet flow rate, m3 /s


S swirl number S D  Dt Xi /2Ai
Xi centroid of the inlet area, mm

Greek Letters
˛ı swirl chamber taper angle
ˇı first taper angle
" separation efficiency
ı second taper angle

Anda mungkin juga menyukai