Вы находитесь на странице: 1из 7

Journal of Petroleum Science and Engineering 157 (2017) 117–123

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Flow of heavy crude oil-in-water emulsions in long capillaries


simulating pipelines
Alexander Ya. Malkin, Maria V. Mironova, Sergey O. Ilyin *
A.V. Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, 29 Leninsky Prospect, 119991 Moscow, Russia

A R T I C L E I N F O A B S T R A C T

Keywords: We studied the rheological properties of heavy oil-in-water emulsions with the final goal of decreasing the high
Heavy crude oil viscosity of oil in pipeline transportation. The emulsions containing 3% silica or 5% montmorillonite were sta-
Oil-in-water emulsions bilized by Triton X-100. All emulsions were viscoplastic media and also demonstrated viscoelasticity. At low shear
Rotation viscometry
stresses, emulsions are in gel-like state that was reflected by the independence of the elastic modulus from fre-
Tube transportation
quency. The yield stress of basic emulsion is rather low (~1 Pa), while the introduction of solid particles leads to a
Wall sliding
remarkable increase of the yield stress. It was shown that direct transition of the rheological data for oil emulsions
measured in a rotational device to capillary rheometer appeared impossible. The apparent viscosity of the samples
estimated in tubes (as a model of a capillary viscometer) was not correlated with the shear viscosity measured on
a rotational rheometer and overall was dependent on the tube/capillary diameter. The reason for these effects
may be related to the stratified flow through a tube consisting of near-wall slippage or the existence at the stream
periphery of a low-viscosity layer. This phenomenon should be taken into account in designing pipe trans-
portation lines when viscometric data are obtained in the other conditions of flow.

1. Introduction points of view. In fact, preparing oil-in-water concentrated emulsions


allowed for decreasing in the viscosity of heavy crude oil by several or-
Today, interest in heavy crude oils is constantly growing due to ders of magnitude (Malkin et al., 2016). Therefore, it is necessary to
gradual exhaustion of standard oil sources and the discovery of huge stress that the evaluation of the emulsion viscosity cited in the majority of
fields of heavy oil around the world, with the latter accounting for no less publications is rather conditional and ambiguous. All oils and oil-based
than two thirds of the explored reserves of hydrocarbons (Alboudwarej emulsions in particular are rheologically complex media. These fluids
et al., 2006; Akram et al., 2014). At the same time, prospects for a rapid demonstrate not only the viscoplastic behavior, yielding effect, or non-
increase in heavy oil production meet serious limitations due to its high Newtonian flow (Foudazi et al., 2015), but primarily the heterogeneity
viscosity. This leads to increased costs of transportation of crude oil and of the flux of multi-component fluids expressed in stratified flow and wall
thereby impairs economic evaluation of the use of heavy oil. To solve this effects, which depends on the geometry of the flow. For example, phe-
issue, various technological solutions for viscosity reduction have been nomena such as the dependence of the rheological properties of waxy
advanced. Among them, the following ideas have been discussed: heating crude oils on the gap between solid surfaces in a measuring rotation
to the temperature at which the viscosity decreases to an acceptable device have been described (Japper-Jaafar et al., 2015). Also, the influ-
level, dilution of heavy oil with light oil to reach the necessary viscosity ence of the pipe diameter on the rheology of such oils has been observed
level of a blend, use of drag reduction additives, arranging a multi-layer (Foissen et al., 2013).
flow by creating a low-viscosity wall layer (annular and core flow), heavy Effects of this kind are not typical for heavy crude oil itself, which
oil deasphalting (Ilyin et al., 2016a), and transformation of crude oil into usually contains a low quantity of paraffin wax. The behavior of these oils
oil-in-water emulsion. One can find the collation of all these methods as is almost like Newtonian viscosity in a wide temperature range (Ilyin
well as numerous references to earlier publications in the review (Mar- et al., 2016b). However, the transformation of crude oil into emulsions
tínez-Palou et al., 2011). that are viscoplastic and thixotropic fluids does not allow the features of
The development of oil-in-water emulsions looks rather attractive the flow mode and the specificity of the method used for measuring the
among the proposed methods from both technological and economic viscosity to be neglected. Moreover, the most important point is the issue

* Corresponding author.
E-mail address: s.o.ilyin@gmail.com (S.O. Ilyin).

http://dx.doi.org/10.1016/j.petrol.2017.07.024
Received 8 December 2016; Received in revised form 28 June 2017; Accepted 11 July 2017
Available online 12 July 2017
0920-4105/© 2017 Elsevier B.V. All rights reserved.
A.Ya. Malkin et al. Journal of Petroleum Science and Engineering 157 (2017) 117–123

2.2. Methods

The rheological properties of crude oil emulsions were measured on a


RheoStress 600 rotational rheometer (Thermo Haake, Germany), a
Rosand RH10 capillary viscometer (Malvern Instruments, UK), and a
home-made pressure controlled laboratory viscometer.
The measuring device was a cone-and-plate pair
(diameter ¼ 60 mm; angle between conical surface and plate ¼ 1
grade). The rotational rheometer was used to measure the properties in
the stationary and oscillation modes of deformation. The viscosity was
measured for a rate-controlled shear rate in the range of
1  104–100 s1. Viscoelastic properties in the linear limit were
measured in the frequency range of 0.05–600 s1 at the same shear
Fig. 1. Principle scheme of a pressure-controlled viscometer: 1 – compressor, 2 – valve, 3
stress amplitude of 1 Pa. The relative error in measuring the rheological
– manometer, 4 – shutoff valve, 5 – barrel with a sample, 6 – tube (capillary), 7 – open properties does not exceed 5%.
container, 8 – analytical balance, 9 – computer. The Rosand RH10 capillary viscometer worked at a preset velocity
and was supplied with a non-standard homemade brass tube with a
of whether we can apply the characterization of the rheological proper- length, l, of 5 m and inner diameter, d, of 0.8 mm. This long tube with l/
ties obtained by the technique of rotational rheometry (commonly used d ¼ 625 was considered as a model of a pipeline where the apparent
for estimating the rheology of crude oils) for predicting and designing the (average) shear rate varied from 10 to 1000 s1. The relative error in the
performance of these systems in pipeline transportation. viscosity measurements does not exceed 5%.
This study was carried out as an attempt to answer this question. Its The second capillary device used as a model of a pipeline worked in
main content was the correlation of data of the rotation rheometry and the pressure-controlled mode. A sketch of it is shown in Fig. 1. The
the behavior of the oil-in-water emulsions during their flow through long viscometer consisted of a barrel, 5, containing the studied sample.
tubes (capillaries), which modeled the pipeline transport. We considered Pressure was created by compressed air supplied from a compressor, 1,
such an emulsion stabilized by a common surfactant as well as these regulated by a valve, 2, and measured by a manometer, 3, with a scale of
materials with the addition of solid micro(or nano)-particles, which can 3 atm (±0.05 atm). A shutoff valve, 4, was used for the start of the
be either natural components of crude oil or stabilizers in Pickering- experiment. After opening the shutoff valve, the fluid under study begins
type emulsions. to flowed through a rigid polypropylene tube (capillary), 6, with a known
length (l ¼ 278 mm) and diameter (d ¼ 3.25 mm) into an open container,
2. Experimental section 7, placed on an analytical balance, 8. The balance (AND EK-12Ki) was
connected through a COM-interface to the programmed computer mod-
2.1. Materials ule 9. This allowed for continuous measurement of the rate of flow. Ex-
periments were carried out in the stress-controlled mode in the range
Heavy crude oil from Yaregskoye oil field (Lukoil, Russia) was used as from 10 to 100 Pa. During the tests, the pressure uniformly acted on the
a representative sample. The density at 20  C was equal to 0.942 g/cm3, surface of a sample due to the large distance (10–15 cm) between the
which corresponded to API gravity ¼ 18.1. According to the classification injection nozzle and the sample. The relative error in the viscosity
of the American Petroleum Institute, this oil belongs to the class of heavy measurements does not exceed 10%.
oils (heavy oils are those with API gravity < 20). The weight fraction of Experimental results were treated using the standard methods
sulfur was 1.24%, and the paraffin wax, asphaltene, and resin fractions (Malkin and Isayev, 2017), namely the shear stress in the rotation de-
were 0.8, 5.7, and 17%, respectively. The viscosity, η, at 20  C was vice was calculated from the torque and the shear rate was calculated
approximately 2 Pa s and this value was constant in the range of shear from the rotational speed. The capillary viscometry data were used for
stress, σ, from 0.1 to 3  103 Pa. calculation of the shear stress at the wall, σ, and shear rate, γ_ , respec-
Oil-in-water emulsions were prepared using distilled water and tively, as
nonionic surfactant Triton X-100 (4-(1,1,3,3-tetramethylbutyl)phenyl-
PR
polyethylene glycol) supplied by AppliChem Panreac. This compound σ¼ (1)
has a rather high hydrophilic–lipophilic balance equal to 13.5. The 2l
concentration of the surfactant in aqueous phase was 5 wt %.
and
Besides oil-in-water emulsions stabilized by this surfactant, addi-
tionally, samples containing micro(nano)-size particles have been stud- 4Q
ied. Two types of solid particles were used: 3 wt % silica (produced by γ_ ¼ ; (2)
πR3
Aldrich) with a specific surface of 395 m2/g and 5 wt % Na-
montmorillonite clay Cloisite Naþ (Southern Clay Products, USA) with where Р is the pressure inside the barrel with the sample (at the capillary
a specific surface of 750 m2/g. The aim of using solid particles was entrance), Q is the volume output measured by the weight method as
twofold: on one hand, they simulated natural components in crude oil, described above, R is the radius of a capillary (long tube), and l is
and on the other hand, they could be an addition surfactant line in its length.
Pickering emulsions (Chevalier and Bolzinger, 2013; Ilyin et al., 2015). No corrections for the non-Newtonian flow and entrance effects were
The model system (direct crude oil-in-water emulsions) contained introduced.
65 vol % oil. The emulsions were prepared by continuous addition of the In discussing the experimental data, the Reynolds number, Re, was
necessary volume of oil to the aqueous solution of surfactant, which also calculated by the ordinary formula:
could also contain solid particles. The dispersion was performed in an
V0 dγ 2Qρ
IKA Ultra-Turrax T10 rotary homogenizer. Mixing was continued for Re ¼ ¼ (3)
10 min at a rotor rotation rate of 30  103 rev/min. The emulsions were ηg πRη
stable: precipitation of the aqueous phase did not take place for at least Using the Hagen-Poiseuille equation for viscosity, this formula can be
several weeks. rearranged and Re is expressed via measured values:

118
A.Ya. Malkin et al. Journal of Petroleum Science and Engineering 157 (2017) 117–123

Fig. 2. Photomicrographs of 65% emulsions without solid particles (a) and emulsions with 3% SiO2 (b) or 5% clay (c).

3. Results and discussion


1:6Q2 lρ
Re ffi (4)
PR5 3.1. Structure of emulsions
In these formulas, V0 is the average linear velocity of the flux corre-
sponding to the volume output, γ is the specific weight, ρ is density, η is Photomicrographs of emulsions are shown in Fig. 2 where one can see
viscosity, and g is the gravitational acceleration. All values used for cal- that droplets in emulsions are not deformed in spite of the high content of
culations were expressed in the SI system of units. disperse phase. The size of droplets can reach 30 μm, but in emulsions
The morphology of the prepared emulsions was observed using a containing solid particles, the size of droplets is smaller and lies mainly in
Biomed 6P optical microscope (Russia). The size distribution of the the range of 1–20 μm.
droplets in the emulsions was found by the method of dynamic light The droplets in emulsions with Cloisite Naþ are smaller than those in
scattering by means of a Zetasizer Nano ZS analyzer (Malvern In- emulsions with SiO2, which can be explained by the twofold larger
struments, UK). Samples for the analysis were prepared by dilution of the specific area of the surface of clay in comparison with silica.
oil emulsion droplets in 5 ml of distilled water. The size distribution of droplets is presented in Fig. 3. As seen,
All experiments were carried out at 21 ± 1  C. droplets with clay are smaller than droplets with silica and are charac-
terized by a narrower distribution.

Fig. 4. Dependencies of the apparent viscosity on shear stress for basic crude oil and 65%
Fig. 3. The size distribution of droplets in emulsions with the addition of solid particles. oil-in-water emulsions.

119
A.Ya. Malkin et al. Journal of Petroleum Science and Engineering 157 (2017) 117–123

3.2. The rheology of emulsions (rotational rheometry)

The basic heavy crude oil is a Newtonian fluid with high viscosity
close to 2 Pa s (Fig. 4, curve 1). Emulsion containing 65% oil is a non-
Newtonian fluid and its viscosity decreases by a factor of ~1000 along
with an increase in the shear stress (Fig. 4, curve 2). The non-linear ef-
fects may be due to the rupture of inter-droplet bonds and the orientation
of droplets in shearing. In the range of high shear stresses, the viscosity of
emulsion appears to be more than an order of magnitude higher in
comparison with basic oil.
As might be expected, emulsions demonstrate obvious viscoplastic
behavior, expressed in the sharp decrease in apparent viscosity in a
narrow stress range. If the emulsion does not contain solid particles, its
yield stress does not exceed 1 Pa (Fig. 4, curve 2), while the introduction
of solid particles results in a strong increase in the yield stress, which rises
to 10 Pa for emulsion with 5% silica (Fig. 4, curve 3) and to ~30 Pa for
emulsion with 5% clay (Fig. 4, curve 4).
Beyond the yield stress when the structure is destroyed (Malkin et al.,
2017), the apparent viscosity of emulsions becomes much lower than the
viscosity of crude oil.

3.3. Viscoelasticity of emulsions (measurements at the limit of very low


deformations)

Basic crude oil demonstrates viscoelastic properties at low tempera-


tures, possibly due to the presence of asphaltenes and other high-
molecular-weight compounds (Ilyin et al., 2016a, 2016b). To an even
greater extent, viscoelasticity is inherent for emulsions; it is only slightly
expressed in emulsion of pure oil (actually the storage modulus is not
measured by experiment) but becomes strongly expressed in emulsions
containing solid particles (Fig. 5).
The most interesting result presented in Fig. 5b is the independence of
the storage modulus from frequency in a rather wide frequency window, in
which the storage modulus is much higher than the loss modulus. Such
behavior is typical for matter in solid-like state and this means that
emulsions at low shear stresses (in the linear domain of viscoelasticity) are
gels that are unable to flow. The spatial structure of emulsions responsible Fig. 5. Dependencies of the storage modulus (filled symbols) and loss modulus (empty
symbols) on frequency for crude oil (1), oil-in-water emulsions without solid particles (2),
for such a state is formed by contacting gels with interface boundary
and emulsions with 3% silica (3) or 5% clay (4).
enriched with adsorbed solid particles. The narrower size distribution of
particles in emulsion with clay provides higher values of the storage
modulus in the same way as higher values of the storage modulus are is also worth mentioning that flow properties measured by the capillary
characteristic for monodisperse polymers in comparison with samples that method demonstrate slightly non-Newtonian behavior, although this is
are the same but polydisperse (Malkin and Isayev, 2017). not clearly noticeable when using a log scale for the viscosity.
As can be seen, the apparent viscosity estimated by the tube flow data
3.4. Viscometry in the flow through long tubes (capillaries)

Oil-in-water emulsions especially those that have been modified by


the addition of solid particles are viscoplastic non-Newtonian fluids and
the flow of such media at high shear stresses can be heterogeneous (or
stratified) (Malkin et al., 2012; Malkin and Kulichikhin, 2015). This can
occur as a well-known effect of shear banding (Møller et al., 2008;
Fielding, 2016; Divoux et al., 2016). Then the relationship between the
pressure gradient and the average velocity in the flow (as the measure of
the apparent viscosity) depends on the velocity profile and is therefore
not an objective characteristic of the rheological properties of a fluid.
So it is crucial to consider the correlation between the objective
rheological characteristic of matter obtained in the homogeneous flow
field (Fig. 4) and the peculiarities of the flow through a long capillary
(tube) as a model of a pipeline.
Appropriate experimental data are presented in Fig. 6 where the filled
symbols reproduce the rotational rheometry data and the open symbols
are the results of capillary viscometry treated by using Eqs. (1) and (2).
Comparison of the two sets of data shows that the flow curves are
Fig. 6. Dependencies of the apparent viscosity on shear stress for crude oil (1), oil-in-
equivalent for crude oil while the results of measurements appear water emulsions without solid particles (2), and emulsions with 3% silica (3) or 5%
different regardless of the level of the yield stress (low yield stress for clay (4). Rotational rheometry data are shown by filled symbols. Capillary viscometry data
pure emulsions and high yield stress for emulsions with solid particles). It (the tube with a diameter of 0.8 mm) are shown by empty symbols.

120
A.Ya. Malkin et al. Journal of Petroleum Science and Engineering 157 (2017) 117–123

the measurements do not coincide with one another because the


apparent viscosity appears higher for the tube of larger diameter. This
observation can be explained if it is supposed that the stratified flux takes
place due to wall effects (either real wall slip or formation of a thin low-
viscosity layer).
Indeed, if wall slip with velocity w happens, the volume output is
calculated as

Q ¼ Q0 þ πR2 w (5)

where the first member in Formula (5) is the volume shear flow in the
bulk of a tube and the second member is due to wall slip of fluid and
movement of the material as a whole with the velocity w. Then the
following relationship is valid:

1 1 4
¼ þ w (6)
ηapp η Rσ

Fig. 7. Dependencies of the apparent viscosity on shear stress for 65% oil-in-water where η is the real viscosity determined by shearing and ηapp is the
emulsion without solid particles (1) and emulsions with 3% silica (2) and 5% clay (3).
apparent viscosity calculated by the Hagen-Poiseuille equation for a
The results have been obtained for tubes with diameters of 0.8 mm (filled symbols) or
3.25 mm (empty symbols). Newtonian fluid. Then one can see that the apparent viscosity ηapp should
increase with increasing tube radius. The numbers on the y-scale in Fig. 7
is at least twofold lower than that measured in the homogeneous con- are just these apparent viscosity values and they approached the real
ditions for the same shear stress. shear viscosity as the radius of the tube increased.
The crucial experiment for understanding the peculiar features in the The graphs in Fig. 8 built using the coordinates of Eq. (6) give some
flow of emulsions is testing long tubes with different diameters. The re- idea of the relationship between the real shear viscosity of emulsions and
sults of such an experiment are presented in Fig. 7 show that the data of the role of wall slip.

Fig. 8. Presentation of the experimental data using the coordinates of Eq. (6) at different shear stresses for emulsion without solid particles (a) and emulsion with 3% silica (b) or 5%
clay (c).

121
A.Ya. Malkin et al. Journal of Petroleum Science and Engineering 157 (2017) 117–123

This relationship shows the physical sense of the inequality, that is,
the relation between the shear rate in the volume of fluid ð_γ Þ and the
apparent velocity gradient (w/R) due to the wall slip.
The following example illustrates the estimation of the tube radius
when the wall slip can be neglected. For oil-in-water emulsion without
solid particles with η ¼ 0.15 Pa s, σ ¼ 10 Pa, and w ¼ 14.5 mm/s, the
strong inequality will be executed for R > 9 mm, which is much larger
than the tube with R ¼ 1.625 mm used for laboratory measurements.
For emulsions without solid particles, the slip velocity increases along
with the shear stress (Fig. 9b), while for emulsions with silica or clay, the
slip velocity passes through maxima. One can suppose that a decrease in
velocity is related to the destruction of the laminar wall-adjacent layer by
solid particles rotating in flux. It should be emphasized that the slip ve-
locity does not depend on the tube diameter but is determined only by
the pressure gradient and the nature of emulsions (i.e. the structure of
near-wall layer).
The obtained experimental results can be treated as the consequence
of stratified flow of emulsions. Indeed, it is well known that such a mode
of movement is typical for multicomponent systems (Malkin et al., 2012).
The stratified flows have been also observed for crude oil-water mixtures
as was discussed in the recent publication based on the high resolution
visualization of the flux (Loh and Premanadhan, 2016). In this study, it
has been shown that the dispersed flow regime is more typical than the
stratified regime for heavy oils in comparison with light oils. Possibly,
this is also true for the heavy oil emulsions. In this case, we meet with a
flow of a homogeneous dispersion (emulsion) with the wall slip rather
than with a real stratified flux.
In conclusion, it is worth stating that all experimental data obtained
in this study were obtained at low Reynolds numbers. Calculations show
that for the flow of heavy crude oil Re << 1 and for the emulsions under
study, Re lies in the range from 2  103 to 40. So, formally, the flow in
this study was always laminar. However, the definition of the Reynolds
number for stratified flows is rather uncertain.

4. Conclusions

The creation of oil-in-water emulsions from heavy crude oil leads to a


Fig. 9. Shear stress dependencies of real viscosity (a) and wall slip velocity (b) for
emulsion without solid particles (1) and emulsion with 3% silica (2) or 5% clay (3). The significant decrease in viscosity at high stresses, which is important for
slip velocities have been calculated for tubes with diameters of 0.8 mm (filled symbols) or the technology of pipeline transportation. Structuring concentrated
3.25 mm (empty symbols). emulsions results in the appearance of viscoplasticity (yield stress) and
viscoelasticity. The presence of solid micro (or nano)-particles in emul-
The viscosity values obtained by extrapolation to 4/Rσ ¼ 0 (i.e. R → ∞) sion, which can be either natural impurities or Pickering stabilizer, in-
allow us to build the real flow curves which are the objective presentation fluences the morphology of emulsions and leads to an increase in the
of the viscous properties of emulsions under study (Fig. 9a). For example yield stress and viscosity. In the domain of low stresses (linear visco-
the viscosity values at shear stress of 10 Pa are equal to 0.15, 0.20, and elasticity), the storage modulus does not depend on frequency, which is
0.27 Pa s for emulsion without solid particles and emulsion with 3% silica typical for emulsions in gel state or solid-like state.
or 5% clay, respectively. The first of these values is in satisfactory agree- The behavior of emulsions in the flow depends on the geometry of
ment with the viscosity at the plateau at stresses above the yield stress deformation. The objective rheological properties of emulsions measured
obtained by the rotational rheometry method. In the other cases, such a in homogeneous shearing cannot be used directly for predicting their
comparison is impossible because the level of plastic viscosity on the behavior in flow through capillaries and designing pipelines for oil
rotational rheometer has not been reached (filled symbols in Fig. 6). The transportation because the apparent viscosity found in the capillary
slip velocities are 14.5, 5.5, and 1.0 mm/s for emulsion without solid viscometer does not coincide with the viscosity measured in a rotational
particles and emulsion with 3% silica or 5% clay, respectively. rheometer. Moreover, the “capillary” apparent viscosity depends on the
These arguments emphasize the limited possibilities for application of diameter of the tube. These effects are evidently due to inhomogeneity of
the rotational rheometry data for designing the flow in pipelines. Eq. (6) the flux in the tube due to wall slip and the appearance of stratified flow.
shows that the wall slip can be neglected if These peculiarities in the behavior of concentrated oil-in-water emul-
sions should be taken into account when applying the results of the
4η rheological tests to real pipelines. The volume shear flow becomes
w< <1 (7)
Rσ dominant only in the flow of tubes of large diameter, while whether the
This means that the slip effect can be insubstantial only for tubes of diameter is large enough depends on the composition of the emulsion.
large diameter. This inequality can be written in the following alterna-
tive form: References

4η 4w Akram, F., Stone, T., Bailey, W., Freeman, M., Law, D., Woiceshyne, G., Yeung, K., 2014.
w¼ < <1 (8) Warming to heavy oil prospects. Oilfield Rev. 26, 4–15.
Rσ R_γ Alboudwarej, H., Felix, J., Taylor, S., Badry, R., Bremner, C., Brough, B., Skeates, C.,
2006. Highlighting heavy oil. Oilfield Rev. 18, 34–53.

122
A.Ya. Malkin et al. Journal of Petroleum Science and Engineering 157 (2017) 117–123

Chevalier, Y., Bolzinger, M.-A., 2013. Emulsions stabilized with solid nanoparticles: Japper-Jaafar, A., Bhaskoro, P.T., Sean, L.L., Sariman, M.Z., Nugroho, H., 2015. Yield
Pickering emulsions. Colloid Surf. A 439, 23–34. stress measurement of gelled waxy crude oil: gap size requirement. J. Newt. Fluid
Divoux, T., Fardin, M.A., Manneville, S., Lerouge, S., 2016. Shear banding of complex Mech. 218, 71–82.
fluids. Annu. Rev. Fluid Mech. 48, 81–103. Loh, W.L., Premanadhan, V.K., 2016. Experimental investigation of viscous oil-water
Fielding, S.M., 2016. Triggers and signatures of shear banding in steady and time- flows in pipeline. J. Petrol. Sci. Eng. 147, 87–97.
dependent flows. J. Rheol. 60, 821–834. Malkin, A.Y., Isayev, A., 2017. Rheology: Concepts, Methods, and Applications, third ed.
Foissen, M., Øyangen, T., Velle, O.J., 2013. Effect of the pipe diameter on the restart ChemTec Publ., Toronto.
pressure of a gelled waxy crude oil. Energy Fuels 27, 3685–3691. Malkin, A.Y., Kulichikhin, V., 2015. Spatial-temporal phenomena in the flows of multi-
Foudazi, R., Qavi, S., Masalova, I., Malkin, A.Y., 2015. Physical chemistry of highly component materials. Appl. Rheol. 25, 35358.
concentrated emulsions. Adv. Colloid Interface Sci. 220, 78–91. Malkin, A., Ilyin, S., Semakov, A., Kulichikhin, V., 2012. Viscoplasticity and stratified flow
Ilyin, S.O., Kulichikhin, V.G., Malkin, A.Y., 2015. Rheological properties of emulsions of colloid suspensions. Soft Matter 8, 2607–2617.
formed by polymer solutions and modified by nanoparticles. Colloid Polym. Sci. 293, Malkin, A.Y., Zadymova, N.M., Skvortsova, Z.N., Traskine, V.Y., Kulichikhin, V.G., 2016.
1647–1654. Formation of concentrated emulsions in heavy oil. Colloid Surf. A 504, 343–349.
Ilyin, S., Arinina, M., Polyakova, M., Bondarenko, G., Konstantinov, I., Kulichikhin, V., Malkin, A., Kulichikhin, V., Ilyin, S., 2017. A modern look on yield stress fluids. Rheol.
Malkin, A., 2016a. Asphaltenes in heavy crude oil: designation, precipitation, Acta 56, 177–188.
solutions, and effects on viscosity. J. Petrol. Sci. Eng. 147, 211–217. Martínez-Palou, R., de Lourdes Mosqueira, M., Zapata-Rend on, B., Mar-Ju
arez, E., Bernal-
Ilyin, S.O., Arinina, M.P., Polyakova, M.Y., Kulichikhin, V.G., Malkin, A.Y., 2016b. Huicochea, C., de la Cruz Clavel-L opez, J., Aburto, J., 2011. Transportation of heavy
Rheological comparison of light and heavy crude oils. Fuel 186, 157–167. and extra-heavy crude oil by pipeline: a review. J. Petrol. Sci. Eng. 75, 274–282.
Møller, P.C.F., Rodts, S., Michels, M.A.J., Bonn, D., 2008. Shear banding and yield stress
in soft glassy materials. Phys. Rev. E 77, 041507.

123

Вам также может понравиться