Sie sind auf Seite 1von 14

SECTION 5

CRUDE DEHYDRATION

CONTENTS
Page No.
1.

INTRODUCTION

2.

EMULSIONS

3.

EMULSION STABILITY

4.

THE CAUSES OF CRUDE OIL EMULSION PROBLEMS

4.1

4.2

5.

PREDICTING DEHYDRATION PROPERTIES


5.1
5.2

6.

Chemical Components
4.1.1 Naturally Occurring Surfactants Resins
4.1.2 Asphaltenes
4.1.3 Wax
4.1.4 Sand
4.1.5 Salt Crystals
4.1.6 Corrosion Products
4.1.7 Other Oilfield Chemicals
4.1.8 Demulsifier Overdosing
Physical Parameters
4.2.1 Temperature and Ageing
4.2.2 Droplet Size Distribution
4.2.3 Viscosity
4.2.4 Low Water Cuts

BP Bottle Test
Pilot Plant Tests

SOLUTIONS TO EMULSION SEPARATION PROBLEMS


6.1

6.2

Engineering
6.1.1 General
6.1.2 Electrostatic Separation
6.1.3 Produced Water Recycling
Demulsifier Chemicals
6.2.1 Composition and Mechanism of Action
6.2.2 Deployment Strategy

7.

THE FUTURE

10

8.

GENERAL READING REFERENCES

11

FIGURES 1- 5

1.

INTRODUCTION

Crude oil is often found in association with water, the oil bearing zone overlying a saline aquifer. As
the reservoir is depleted, a stage is reached (days or years after production start up) when this
formation water is coproduced with the oil. If water injection is employed (to maintain reservoir
pressure) then injection water (usually sea water) may also be coproduced later in the life of the field.
Although water and crude oil are immiscible, the produced fluids are readily emulsified by the
simultaneous action of shear and pressure drop such as occur at chokes or across valves. The
produced emulsions can be stable, i.e. the oil and water phases do not separate readily within the
residence time of separation equipment. Offshore, this is usually a period of minutes (e.g. Forties ca
3 minutes; Magnus, ca. 10 minutes; Buchan ca. 15 to 20 minutes). Even onshore, normal fluid
throughput may not allow sufficient time for dehydration (e.g. Endicott, first stage oil residence time
ca. 9 minutes). In other words, produced emulsions often have to be treated (usually with a chemical
emulsion breaker also called a 'de-mulsifier' or 'demulsifier') to effect efficient dehydration.
The operation consequences of not removing this water would be:
1.
2.
3.
4.
5.

Export oil specifications would not be met.


The water with its associated salt would result in increased corrosion of pipework, pumps and
downstream production equipment.
Unnecessary expense would be involved in pumping or transporting a valueless product by
pipeline or tanker.
Water would occupy vital oil capacity in storage and transport facilities.
Crude refining costs would be higher and therefore the crude would be of lower value.

Hence, for example, the spec for the maximum residual water content of the Forties Main Oil Line is
2%. The TAPS line specification is 0.35%.
Demulsifiers greatly facilitate oil-water separation. However, they cannot perform miracles! It is
therefore a prerequisite when evaluating new field designs, that the facilities proposed for dehydration
(e.g. number, size and type of separators, operating temperatures, sites for chemical injection) are
established as satisfactory - neither under nor over design is economically acceptable.
The presence of high levels of salt in crude exported to refineries is undesirable since, in addition to
corrosion, salt causes saline fouling in refinery heat exchangers and even in trace (ppm) quantities can
poison refinery catalysts. Table 1 lists the maximum amounts of residual water after dehydration that
meet a typical salt target of 25 pounds per thousand barrels (ptb) (1 ptb is approximately 3mg/l).
Table 1
Field
Magnus
Prudhoe
Endicott
Forties
Buchan
Ula

Water (%) to give 25 ptb


0.4
0.3
0.23
0.08
0.04
0.03

Formation Water
Total Salinity (mg/l)
20,000
24,710
33,140
99,000
196,000
253,000

Consequently, even after reducing residual water contents of produced crude oil to achieve maximum
water specifications, further dehydration may be needed to achieve salt specification.
Desalting is usually achieved by "washing" the crude oil with water of low salinity (typically 0.5-5%
by volume). The wash water is usually fresh (or desalinated) water, although sea water is also widely
used. The water is mixed with the crude oil so that dispersed water contact residual droplets of
formation water or salt crystals and the salt moves into the lower ionic strength wash water. At the
same time the extra water helps to destabilise the emulsion which is then "resolved" ("broken" or
"separated") under gravity or in an electrostatic field (e.g. at Das Island and in Northern Kuwait).
Further desalting is performed at refineries, normally in electrostatic separators operating at high
temperatures (100-140C), where residual salt contents are further reduced from ca. 25 ptb to ca.
2ptb. An alternating electric current when applied to low water cut emulsions causes increased
water droplet coalescence and as the water drops out it takes the unwanted salt with it. (In recent
years, AC has been replaced in some cases by pulsed DC, which is beneficial under certain
circumstances).
For reasons similar to those described above, emulsion breaking technology is also applied to the
selection of demulsifiers for wash water desalting and to design of the minimum facilities required to
produce acceptable desalting.

2.

EMULSIONS

Definition - An emulsion is a heterogeneous system consisting of at least one immiscible liquid


intimately dispersed in another in the form of droplets.
Terminology - The phase present as droplets is termed the dispersed phase. The liquid phase
forming the matrix within which these droplets are suspended is called the continuous or external
phase.
In oilfield operations the two common types of emulsion are:
water-in-oil (w/o) and
oil-in-water (o/w) which is observed as a function of the produced water cuts.

3.

EMULSION STABILITY

The dispersed phase consists of microscopic droplets, usually within the size range 0.1 to 100
microns. When two separate liquid phases are converted into an emulsion a very large amount of
new surface area is created. An energy input is required to produce this interfacial area. Therefore
when drops coalesce the free energy of the system is reduced. Thus crude oil emulsions are
thermodynamically unstable. Unfortunately the kinetics of the separation process may be too slow
for our liking.
In discussing emulsion stability it is necessary to distinguish between the mechanisms involved in the
stages of "flocculation" (or "coagulation"), "creaming" and "breaking". (Figure 1).
3.1 Flocculation is the association of droplets, forming three-dimensional clusters, without
coalescence which would form larger individual droplets.

3.2 Creaming is the term given to the fall (w/o emulsions) or the rise (o/w emulsions) of
dispersed droplets under the action of gravity. Clearly, creaming can only arise if the dispersed and
continuous phases are unequal in density. The rate of creaming for a single drop is governed by
Stokes' Law:
V=

2gr2 (Dp)
9z

Where
V

velocity of fall (or rise) of droplets

gravitational constant

droplet radius

Dp

difference in density between the dispersed and continuous phases

viscosity of the continuous phase

3.3 Breaking of an emulsion occurs when droplets which are flocculated or which collide
actually coalesce, i.e. break their individual interfacial films, forming a larger droplet.
It is useful at this point to introduce the concept of interfacial tension. Two immiscible liquids in
contact, e.g. oil and water, possess a tension in the planar interface between them. When an emulsion
is created, curvature is introduced into this interface thus increasing the energy of the system. If the
interfacial tension is low the curvature is more easily formed. Thus smaller droplets can be formed.
(The limiting case is for completely miscible liquids when the tension would be zero.)
The discussion to date has been in terms of two components i.e. oil and water. It is essential that a
third chemical component is present to confer stability to the produced emulsion. This component
will either form a physical (repulsive) barrier to droplet-droplet coalescence, or reduce the interfacial
tension or both. This is discussed further below in section 4.1.

4.

THE CAUSES OF CRUDE OIL EMULSION PROBLEMS

Except where water cuts are extremely high (typically >60-80%), crude oil emulsions are oilcontinuous i.e. of the water-in-oil (w/o) type. The treatment of the reverse oil-in-water type is similar
to that of the oily effluent (see Water Treatment section) and will not be discussed in this section.
Even when water cuts are very high, produced fluids entering the separation equipment usually
consist of some free water and an oil continuous emulsion with the remainder of the water.
If untreated, w/o produced emulsions are often stable with respect to water separation and under
ambient conditions may remain unchanged for several years. In the first instance, therefore, we need
to consider what factors (chemical and physical) give rise to this stability.
4.1

Chemical Components

We can consider these to form two groups, those which lower interfacial tension and those which
form a physical barrier to droplet coalescence. The former facilitate the creation of the emulsion in
the first instance but it is the latter class that is responsible for the main operational problems.
4.1.1

Naturally Occurring Surfactants (Resins)

These are relatively low molecular weight (,2,000) organic species often with acid or amine
functional groups. They have a natural affinity for the oil/water interface to which they can move
rapidly and are primarily responsible for lowering interfacial tension.
4.1.2

Asphaltenes

These are higher molecular weight species and as a result tend to move only slowly to the surface of a
water droplet. However they possess several polar groups per molecule and can therefore attach to
the interface at a number of points. This makes them particularly difficult to displace. Furthermore,
their solubility in crude can be dramatically affected by small changes of temperature, pressure and
loss of light ends. This can result in "solidification" at the interface and the formation of a
particularly effective barrier to coalescence.
4.1.3

Wax

Wax crystals can act in the same way as solidified asphaltenes. In their natural state, they are less
polar than asphaltenes and thus are not strongly attracted chemically to the droplet surface. However,
they can interact with resins or asphaltenes which modify their surface properties and can act as a link
to hold the wax crystals to the interface.
Waxes are of particular relevance to North Sea operations and a study has been carried out at Ninian.
Removal of wax, for example, from (Ninian) crude oil at 35C and 20C does not appreciably change
the density or viscosity of the oil, nor the interfacial tension, but significantly reduces emulsion
stability. (see Figure 2). The wax is thought to be interfering with the transition from flocculation to
coalescence by providing a physical barrier between the droplets.
4.1.4

Sand

Silica is generally too polar and too dense to stabilise produced emulsions. However, formation
sands also contain a percentage of clays. These can either interact with resins to form surface active
aggregates or form a network within the water droplet matrix which acts as a barrier to coalescence.
(NB Like asphaltenes, sand can accumulate in separators thus interfering with emulsion resolution).
4.1.5

Salt Crystals

These are similar to sand in their effect but are of lower density. Some formation waters are so
highly saline that comparatively small changes in temperature can cause precipitation of salt crystals
which then act as a barrier to droplet coalescence.
4.1.6

Corrosion Products

Iron oxides, hydroxides and sulphides can form as a result of highly saline or sour reservoir fluids
passing through wellheads, pipelines, control valves etc. Their mechanism in emulsion stabilisation
is similar to that of salt crystals.
4.1.7

Other Oilfield Chemicals

These can increase emulsion stability in two ways. Firstly, they are often surface active in
themselves because they are required to coat for instance a wax surface (wax inhibitors) or a metal
surface (corrosion inhibitors). Thus, they can be naturally attracted to oil/water interfaces.
Secondly, in some cases demulsifiers can interact with other oilfield chemicals. Corrosion inhibitors
in particular can combine with demulsifiers to create a surface active cocktail which increases

emulsion stability. This phenomenon emphasises the importance of an integrated deployment


strategy for oilfield chemicals.
4.1.8

Demulsifier Overdosing

After an appropriate chemical has been selected for a given crude plant upsets can occur due to
changes in the physical conditions of the fluids or any of the other effects described above. Often the
initial response is to increase the demulsifier dose rate. As demulsifiers are surface active they may
saturate the oil/water interface and can actually make the situation worse. If this occurs the reverse
strategy of decreasing the dose rate should be employed before further increases are made.
4.2

Physical Parameters
4.2.1

Temperature and Ageing

Increasing temperature
-

reduces the bulk viscosity (increasing rate of creaming - Stoke's Law)


increases the frequency of droplet-droplet collisions
facilitates film drainage between droplets
may increase the density difference between oil and water (increasing rate of creaming - Stoke's
Law)

Hence, the overall effect of increasing temperature is to increase the rate of dehydration of crude
emulsions.
In modern field developments where fluids from a large number of wells (some remote) are processed
at the same gathering station (or platform) fluid temperature effects may be crucial. Even if the
temperatures of cool streams are raised during processing, these fluids may have been cold for a
period long enough to "fix" components such as asphaltenes much more firmly at the oil/water
interface, greatly contributing to the difficulties of dehydration.
On the other hand the effects of wax crystal formation can usually be reversed by heating above a
critical temperature, often around 50C.
4.2.2

Droplet Size Distribution

The rate of separation of water from produced crude increases in proportion to the square of the water
droplet radius (Stokes' Law). The droplet size distribution of produced crude oil emulsions is
polydisperse (as against monodisperse). In most cases, >90% by volume of the water is dispersed as
droplets in the size range 5-25 microns, but many (by number) may be smaller than 5 microns. Very
small droplets like these will sediment only slowly under gravity due to thermally driven motions
within the fluids and the viscosity of the bulk phase itself. The droplet size distribution in produced
emulsions is determined by the degree of shearing and mixing. Although crude dehydration is
critically dependent on water droplet size, in reality there may be little opportunity to exercise control
over it.
4.2.3

Viscosity

The rate of separation decreases with increasing crude oil viscosity. Crude oil viscosities can vary
widely, typically (at 15C) from 10 mPas (light, conventional crude oils) to 10,000,000 mPas (heavy
oils, bitumens, tar sand oils). At the appropriate temperatures and shear rates of production, most
crude oils of interest to BP have viscosities in the range 5-100 mPas. Prudhoe and Endicott crude

oils also lie within this viscosity range. However, this average is tending to rise with an increasing
trend towards development of medium gravity/viscosity crude fields.
It is important to realise that the dispersion of water as an emulsion in crude oil can substantially
change the rheological properties of the crude. Viscosity of an emulsion can be many times that of
the dry crude and the viscosity-shear rate properties of the produced fluids may be substantially
changed. In many models of emulsion behaviour and separation processes it would be more accurate
to use the emulsion viscosity rather than that of the crude. The viscosity increase depends upon the
temperature of the emulsion, the shear rate applied to the emulsion, and most importantly, the droplet
size and phase volume (i.e. 'water cut') of the emulsion.
Thus, particularly at low rates of shear, the viscosities of crude oil emulsions can be several orders of
magnitude greater than that of the dry crude oil. The high viscosity impedes sedimentation and can at
low temperatures, low shear, and high water cuts (e.g. >40%), pose a serious emulsion breaking
problem.
4.2.4

Low Water Cuts

Emulsion dehydration ultimately depends on causing water droplet coalescence. This in turn depends
on the number of collisions, which will inevitably be less if there are fewer water droplets present.
Emulsion resolution at water cuts less than 10% can be particularly problematical. Thus
operationally, the most severe problems encountered can be shortly after water breakthrough when
relatively small volumes of water are processed. These may disappear as water cuts rise, but as
discussed elsewhere in this course, this causes different problems.

5.

PREDICTING DEHYDRATION PROPERTIES

The previous section underlines the wide range of interacting parameters which can influence the
stability of crude oil emulsions. No two crudes possess the same chemical or physical characteristics
nor indeed respond to treatment in exactly the same way. It is therefore vital to consider fluid
properties in addition to process design when optimising separation.
Although laboratory studies have revealed, for instance, a broad correlation between crude asphaltene
content and emulsion stability, there is no substitute for carrying out experiments on the particular
crude itself. Indeed, when selecting appropriate demulsifiers this becomes vital.
Two types of tests are conducted within the oil industry for evaluating demulsifier efficiency and
consequently predicting emulsion stability. These are:
(i)

Laboratory Bottle Test

(ii)

Pilot Plant Tests

5.1

BP Bottle Test

The method entails:


(1)

the batch preparation of simulated water-in-crude oil emulsions at the required


temperature

(2)

the addition and dispersion of chemical demulsifier

(3)

Advantages

monitoring (versus time) at the desired resolution temperature, the rate of water
separation, the oil-water interface and separated water qualities. The volumes of
separated water are measured either visually in graduated glass bottles, or through the
remaining water content of the oil phase as determined by Karl Fischer.
-

Disadvantages -

This method allows rapid screening of demulsifiers under a wide range of


conditions.
Small sample requirement
These are static tests
They only provide relative, rather than absolute, performance data
Temperatures above 80C cannot be easily examined
It is difficult to use this test for water cuts of less than 10%.

BP have found notable discrepancies in this separation of water from some crude emulsions when
there has been some delay between sampling and examination. This test can however be adopted for
use at the production wellhead where fresh samples are available, and indeed this is the preferred use
of such a test. However, the often minimal volumes of appraisal and exploration well samples
available often mean that laboratory determinations are inevitable and only possible after completion
of other studies such as PVT or geochemical. The best indication of demulsifier performance is from
bottle test conducted in the field. Laboratory testing, at best, can rank a large number of demulsifiers
prior to field tests.
5.2

Pilot Plant Tests

BP has developed a test rig for studying crudes of 19 API and above. It is highly flexible and can be
adapted to mimic most oil production or refinery dehydration/desalting processes, allowing
examination of the effects of e.g. temperature, water cut or demulsifier.
Emulsions, of dispersed produced water or wash water, are prepared continuously and can be
resolved at temperatures in the range of 0-130C in either gravity or electrostatic separators, where
fluid residence times can be adjusted as required to mimic various production rates. The rig can also
model production from remote facilities with associated emulsion cooling during pipelining.
Nitrogen may be injected to simulate higher GOR's.
Demulsifiers can be injected at a variety of locations into the oil, water and emulsified phases - thus
any oilfield additive can be injected to investigate its 'compatibility' with the demulsifier.
Residual water contents in separator oil are determined by Karl-Fischer analysis and residual salt
contents by conductivity. Separated water quality is measured by solvent extraction and infra-red
(total-oil), filtration or centrifugation (soluble oil) and/or turbidity (clarity).
Advantages

Disadvantages -

Dynamic flow (as against static conditions in the bottle test)


Continuous emulsion resolution and preparation (closely resembles
production/process facilities)
Quantitative assessment of oil and water quality; absolute data corresponds
more closely to those obtained in the field.
High crude oil requirement (several hundred litres)
Test duration (only one, possibly two variables investigated per day)

Despite the usefulness of these test procedures in indicating the trend of emulsion characteristics for a
given crude, production conditions with rapidly flowing live crude are very different. There is no
substitute for field testing, especially for fine chemical selection and dose rate optimisations.

Nevertheless initial laboratory and/or pilot rig studies give a valuable guide to likely conditions,
narrow the range of chemical types to try in field tests and provide wide ranging data for separator
facilities design.
They can also indicate a suitable contingency demulsifier to cope with initial separation problems
when significant water breakthrough occurs, thus allowing service companies time to select the most
cost effective chemical in situ.

6.

SOLUTIONS TO EMULSION SEPARATION PROBLEMS


6.1

Engineering
6.1.1

General

However good a chemical treatment is, it needs time to work. For best results, residence times and
high temperatures are needed.
For economic and space reasons, the size of oil/water separation facilities may be constrained.
Usually it is most cost-effective to use the natural heat from the wellstreams to assist separation.
However, in some cases additional heating may also be required.
Where space allows and where the production is sufficient for more than one train, routing of difficult
fluids to the most appropriate separators may be a possible solution to optimising separation.
6.1.2

Electrostatic Separation

Electrostatic separation involves the application of AC or pulsed DC electric fields of high intensity
to the w/o emulsion. As a result water droplets are polarised (Figure 3) and form chains like strings
of beads within the separator. This greatly increases the opportunity for droplet collision and thus
coalescence.
Electrostatic separators are very efficient and are extensively used in refineries where water contents
in the crude are already low and further desalting is necessary. They may be the only efficient
method of dehydration/desalting for water contents less than 10% although their efficiency is
substantially reduced at water cuts of less than about 5%. Incoming streams are usually mixed with
approx. 5% by volume of a low salinity wash water which increases dehydration and desalting
efficiency by improving droplet "chain" formation.
Such separators are however expensive to operate. At production facilities offshore they are
generally used only for difficult crudes. Elsewhere, they are employed as a final "polishing" to meet
a pipeline specification. This is particularly important to achieve the TAPS specification of only
0.35% water in crude.
6.1.3

Produced Water Recycling

For the reasons outlined above, achieving the degree of droplet contact required for efficient
dehydration may be harder early in a field's lifetime. Shortly after water breakthrough, when water
levels are 5-10%, dehydration can be improved by artificially increasing the water cut by recycling
some of the separated water. At higher water cuts there is likely to be insufficient surface active
material in the crude oil to cover the oil/water interface produced. Very efficient droplet coalescence
results and water specification targets are easier to achieve. It may still be necessary to employ
demulsifier chemical and the level should be adjusted to avoid overdosing or decreasing effluent
water quality.
The increasing complexity of new field developments together with the need to treat the fluids from
several fields at a single point means that facilities design and treatment flexibility are going to be the
keys to success in the future.
6.2

Demulsifier Chemicals

The addition of a properly selected chemical demulsifier at very low concentrations (usually
1-50ppm) can cause rapid and virtually complete breakdown of otherwise stable water-in-crude
emulsions. While elevated temperature and/or an electric field may also be essential, the fact remains
that a demulsifier is still usually required.
6.2.1

Composition and Mechanism of Action

Chemical demulsifiers comprise typically:


40-60% of mixed low molecular weight surface active polymers, usually
nonionic, e.g. ethoxylated phenols
and

60-40% solvent (usually aromatic)

The "surfactant-like" active ingredients contain both oil soluble and water soluble parts within the
same molecule: hence their affinity for oil/water interfaces.
At present the number of formulations, patented and commercially available, is enormous. Tens of
thousands of different demulsifier formulations exist. Indeed, in treating an emulsion in the field, a
demulsifier supplier (service company) may blend the optimum chemical in situ from a range of
primary ingredients.
However, as explained in section 4, the barrier to coalescence of water droplets in crude oil emulsions
is largely physical, comprising a film of resins, asphaltenes, waxes and other solids surrounding the
droplets. The primary function of the demulsifier is to break or displace the film of natural
surfactants which prevent coalescence. This is illustrated in Figure 4.
The surface active behaviour of demulsifier chemicals allows them to replace the natural crude oil
components that stabilise emulsions. The "wrong" chemical can however itself stabilise emulsions.
For instance, surfactants which are too "oil soluble" can stabilise water-in-oil emulsions, whereas a
molecule dominated by 'water soluble' functional groups could stabilise oil-in-water emulsions. The
important point is that there is a region of instability, A, where neither o/w or w/o emulsions are
stabilised by members of a given series of surfactants. Such surfactants are potential demulsifiers
(Figure 5).
Demulsifier formulations may also contain components which have additional functions. These
include "wetting agents" which can act to displace solids away from the interface into the water phase
and "flocculators" which assist the aggregation of water droplets.

6.2.2

Deployment Strategy

Because the mechanism of chemical demulsification involves displacement of indigenous crude oil
surfactants from the oil/water interface, the earlier they can begin this task the better. Injection at the
wellhead just downstream of the chokes i.e. just after the emulsion has formed may be advantageous.
This can overcome problems of ageing and/or low temperatures in streams from remote wells.
Economic constraints may mean that this strategy may only be viable for streams which are
particularly difficult to treat. Additionally, separated free water in subsea multiphase flowlines can
impact corrosion control.

7.

THE FUTURE

The oil industry is examining several new technologies to minimise life cycle costs of developments
by examining smarter methods of dehydrating crudes. With the current need to lift the water that
the reservoir produces along with the oil, and to separate it at the surface and to dispose of the
separated water there is a finite cost associated with producing and processing water. One option
under consideration is to reduce the volume of water that needs to be handled at surface through
partial processing of produced streams. Removing water downhole eliminates the requirement for
any surface processing. Current technology does not go as far as removing all of the water downhole.
However, hydrocyclones have been developed that can remove up to about 80% of bulk water
downhole where it can be re-injected via an Electrical Submersible Pump (ESP) to either a disposal
zone or into the aquifer. For this single stage hydrocyclone to work, the well fluids need to be water
continuous, typically this means water cuts in excess of 80%. Forties are in the process of designing
and installing a downhole oil/water separation system in conjunction with an ESP. A Joint Industry
Programme is examining techniques that could extend this process to wells cutting only 40% water,
for instance by using pre-separation and deoiling hydrocyclones in series, or installing a downhole
centrifuge.
Free Water Knock Out (FWKO) vessels are also being designed for location at remote well pads.
These remove the majority of produced water (around 80%) and reinject it locally, freeing up space
in the flowlines for more oil, reducing chemical usage and also reducing the requirement for water
injection lines. They work on the principle of gravity phase separation, the same as 3-phase
separators in the main processing facilities. Wytch Farm are in the process of installing such a vessel
at their M site.
A further Joint Industry Project to design and construct a Configurable Subsea Water Separation
System (CoSWaSS) has been examining water knock out and local re-injection at the seabed. Norsk
Hydro announced that they were sufficiently convinced in the technology to install a water separator
on Troll C. This Project are now evaluating alternative compact partial processing technologies with
the aim to prove operation of various components and systems not part of the Troll trial.

10

8.

REFERENCES

The following represents a general reference list for emulsion properties, resolution and oilfield
operations.
1.

Characterisation of Crude Oil-in-Water Emuslions, pp165-181, Journal of Petroleum Science &


Engineering, Peter E Clark & Ali Pilehvari

2.

Chemical Resolution of Petroleum Emulsions. pp 46-66 Louis T Monson

3.

Designing and Selecting Demulsifiers for Optimum Performance based on Production Fluid
Characterisation. SPE 16285, P D Berger, C Hsu, J P Arendell

4.

Droplet Size Analysis: A New Tool for Improving Oilfield Separations, SPE 18204, D A
Flanigan

5.

Effect of pH on Interfacial Films and Stability of Crude Oil/Water Emulsions, JPT, March
1968, pp 305-312, J E Strassner

6.

Factors Influencing the Stability of Water-In-Crude Oil Emulsions: A Literature and


Experimental Study, BP Sunbury Report CSB/138 463, S E Taylor

7.

Improved Demulsifier Chemistry: A Novel Approach in the Dehydration of Crude Oils, SPE
18481, F Staiss, R Bohm, R Kupfer

8.

Interfacial Activities and Porphyrin Contents of Petroleum Extracts, pp 1759-1765, Industrial


& Engineering Chemistry, Aug 1953, Vol 45, No 8, H N Dunning, J W Moore

9.

Investigations into the Electrical and Coalescence Behaviour of Water-In-Crude Emulsions in


High Voltage Gradients. Colloids & Surfaces pp 29-51, Elsevier Science Publishers,
Amsterdam, S E Taylor

10.

Oil Field Emulsions & Their Electrical Resolution. 20pp L C Waterman, R L Pettefer

11.

Optical Studies of Coalescence in Crude Oil Emulsions, pp 1-8 J Pet Sc & Eng, 9 (1993), D J
Miller, R Bohm

12.

Overview of Emulsion Stability, pp 243-256, P Walstra

13.

The pH-Dependent Colloidal Stability of Aqueous Montmorillonite Suspensions, pp 71-80,


Colloids and Surfaces, Elsevier Science Publishers, Amsterdam 1990, E Tombacz, I Abraham,
M Gilde, F Szanto

14.

A Quantiative Characterisation of pH-Dependent Systems, pp35-44, Paul Cratin, Industrial &


Engineering Chemistry, Vol 2, Feb 1969

15.

Resolving Crude Oil Emulsions, pp 771-772, Chemistry & Industry, Oct 19, 1992, S E Taylor

16.

Spreading Coefficients & Hydrophile Lipophile Balance of Aqueous Solutions of Emulsifying


Agents. pp 1681-1683 AMR, Vol 63, Oct 1959, S Ross, E S Chen

17.

Water-In-Crude Oil Emulsions from the Norwegian Continental Shelf, Part 2, Chemical
Destabilisation and Interfacial Tensions, pp 389-398, Colloid & Polymer Science, Vol 268, J
Sjoblom, H Soderlund, S Lindblad, E J Johansen, I M Skjarvo

18.

Water-In-Crude Oil Emulsion Stability & Emulsion Destabilisation by Chemical Demulsifiers,


pp 100-108, JCPT, New Technology, Apr-Jun, 1968, Montreal, T J Jones, E L Neustadter, K P
Whittingham

11

Processes Involved in Emulsion Instability

Figure 1

Breaking

Sedimentation
(Creaming)
Original Dispersion
Flocculation

Das könnte Ihnen auch gefallen