You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/251560802

Wax formation in oil pipelines: A critical review

Article  in  International Journal of Multiphase Flow · September 2011


DOI: 10.1016/j.ijmultiphaseflow.2011.02.007

CITATIONS READS

243 4,389

4 authors, including:

Dhurjati Prasad Chakrabarti Angelus Pilgrim


University of the West Indies, St. Augustine University of the West Indies, St. Augustine
47 PUBLICATIONS   625 CITATIONS    7 PUBLICATIONS   269 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Industry 4.0 View project

Symbiotic Systems View project

All content following this page was uploaded by M. K. S. Sastry on 04 May 2020.

The user has requested enhancement of the downloaded file.


International Journal of Multiphase Flow 37 (2011) 671–694

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / i j m u l fl o w

Review

Wax formation in oil pipelines: A critical review


Ararimeh Aiyejina a, Dhurjati Prasad Chakrabarti a,⇑, Angelus Pilgrim a, M.K.S. Sastry b
a
Department of Chemical Engineering, The University of the West Indies, Trinidad and Tobago
b
Department of Electrical and Computer Engineering, The University of the West Indies, Trinidad and Tobago

a r t i c l e i n f o a b s t r a c t

Article history: The gelling of waxy crudes and the deposition of wax on the inner walls of subsea crude oil pipelines
Received 23 December 2010 present a costly problem in the production and transportation of oil. The timely removal of deposited
Received in revised form 9 February 2011 wax is required to address the reduction in flow rate that it causes, as well as to avoid the eventual loss
Accepted 20 February 2011
of a pipeline in the event that it becomes completely clogged. In order to understand this problem and
Available online 27 February 2011
address it, significant research has been done on the mechanisms governing wax deposition in pipelines
in order to model the process. Furthermore, methods of inhibiting the formation of wax on pipeline
Keywords:
walls and of removing accumulated wax have been studied to find the most efficient and cost-effective
Waxy crude oil
Oil-pipe
means of maintaining pipelines prone to wax deposition. This paper seeks to review the current state of
Solid–solid transition research into these areas, highlighting what is so far understood about the mechanisms guiding this
Solid–liquid equilibrium wax deposition, and how this knowledge can be applied to modelling and providing solutions to this
Wax precipitation problem.
wax removal Ó 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
2. Detection of deposited wax . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
2.1. Detecting blockages. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
2.2. Detecting wax deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
3. Wax deposition mechanisms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
3.1. Molecular diffusion mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
3.2. Soret diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
3.3. Brownian diffusion mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
3.4. Gravity settling mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
3.5. Shear dispersion mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
3.6. Shear stripping mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 674
3.7. Nucleation and gelation kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 674
3.8. Deposition in two-phase flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 674
4. Effect of emulsified water on gelation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675
5. Cloud point, pour point and gel point correlations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675
6. Review of some existing wax deposition models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675
6.1. Thermodynamic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 676
6.2. Hydrodynamic model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 677
7. Wax aging models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679
7.1. Counter diffusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679
7.2. Ostwald ripening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 680
8. Correct analogies for correlated heat and mass transfer in turbulent flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 680
9. Inhibition of wax deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 681

⇑ Corresponding author. Address: Dept. of Chemical Engineering, The University of The West Indies, St. Augustine, Trinidad and Tobago. Tel.: +1 868 6622002x4001; fax: +1
868 6624414.
E-mail address: dhurjatiprasad@yahoo.co.in (D.P. Chakrabarti).

0301-9322/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijmultiphaseflow.2011.02.007
672 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

9.1.
Chemical inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 681
9.2.
Types of chemical inhibitors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 682
9.2.1. Ethylene copolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 682
9.2.2. Comb polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 682
9.2.3. Wax dispersants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
9.2.4. Polar crude fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
9.2.5. Short-chain alkanes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
9.3. Surfaces that prevent wax deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686
9.4. Cold flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686
10. Wax removal methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
10.1. Pigging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
10.2. Inductive heating. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
10.3. Biological treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
11. Restart of gelled pipelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 688
11.1. Time-dependent gel degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 688
11.2. Examples of restart models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
12. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692

1. Introduction oping methods to mitigate and treat the formation of paraffin


layers in pipelines.
Wax build-up is a complex and very costly problem for the
petroleum industry, widely reported and studied by researchers
in decades past (Reistle, 1928, 1932; Bilderback and McDougall, 2. Detection of deposited wax
1963; Haq, 1978). For subsea pipelines, in particular, it has become
especially important to solve the issue of wax build-up, as large- 2.1. Detecting blockages
scale oil production in colder regions will be faced with more
severe wax precipitation (Smith and Ramsden, 1978; Asperger In order to experimentally explore wax deposition in the field or
et al., 1981). to determine the locations of particularly large wax deposits or
Wax precipitation within pipelines at and below the Cloud even complete plugs, methods are needed for detecting the extent
Point or Wax Appearance Temperature (WAT) can lead to gelling of wax deposition at different points in a pipeline or of detecting
that inhibits flow by causing significant non-Newtonian behaviour the location of plugs. Pressure echo techniques can be used to find
and increasing effective viscosities as the temperature of a waxy the location of a blockage by measuring the time for a pressure
crude oil approaches its Pour Point (Pedersen and Rønningsen, wave to be reflected back along the pipeline from the point of
2003). Alternatively, when just the pipeline wall is below the blockage (Chen et al., 2007). Alternatively, the pipeline could be
WAT, this promotes the deposition of a layer of paraffin molecules pressurized and then a special tool with a calliper and video cam-
that can grow over time, constricting flow. This is especially prob- era on a remotely-operated submersible could be used to measure
lematic for pipelines in deep-sea environments, as, even in rela- the external diameter of the pipeline. Upstream of the blockage,
tively warm climates, the water temperature will be on the order but not downstream of it, an appreciable difference in the diameter
of 5 °C (Azevedo and Teixeira, 2003). can be detected when the pipeline is pressurized (Sarmento et al.,
Some researchers, such as Carmen García et al. (2001) and 2004).
Carmen García and Urbina (2003), have studied correlations be-
tween the properties of crude oils and their flowing properties,
including the precipitation and deposition of wax during flow. 2.2. Detecting wax deposits
Models have been developed to predict the onset of wax precipita-
tion and the deposition of wax along pipeline walls. However, Traditional experimental methods for measuring the extent of
accurately modelling deposition in pipelines can be a complex wax deposits include direct methods such as pigging and the
and difficult undertaking, because, while precipitation is mainly a ‘‘take-out’’ method, in which a section of pipe is removed and
function of thermodynamic variables such as composition, the volume of wax inside measured. Additionally, pressure drop
pressure and temperature, deposition is also dependent on flow and heat transfer methods can be used to measure wax deposits
hydrodynamics, heat and mass transfer, and solid–solid and sur- indirectly without down time (Chen et al., 1997). Zaman et al.
face–solid interactions (Hammami et al., 2003). Only recently has (2004) explored alternative methods of measuring wax deposition
a model been developed that incorporates correct analogies for in pipelines. Firstly, they experimented with measuring light
heat and mass transfer. absorption through crude oil using a light source and a detector
This paper reviews cases where researchers have studied circuit mounted within a pipe. They found that, in laboratory tests,
ways to model wax deposition and the aging of wax deposits this detector circuit proved capable of detecting contamination
in pipelines; methods of measuring wax build-up in pipelines; even with a very small percentage present. The use of ultrasound
methods of inhibiting this deposition; wax removal methods; for solid detection, also explored by Zaman et al. (2004) proved
and restart procedures for pipelines gelled with waxy crude. In very successful in detecting extremely small solid grains. Finally,
doing so, this paper, as one goal, seeks to show how our under- they were able to use a strain gauge to detect very small changes
standing of these mechanisms has developed, to highlight areas in pipeline weight associated with wax deposition. However, all
where further understanding of these mechanisms is still needed, of these methods were only tested with small-scale laboratory rep-
and to show how well our current correlations can be applied to resentations of actual systems. Practical methods for application of
the accurate prediction of wax deposition. Furthermore, this pa- these tools to actual subsea pipelines would still need to be de-
per seeks to highlight the progress that has been made in devel- signed. Zaman et al. (2006) have also experimented with the use
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 673

of a laser spectroscope to detect paraffin in paraffin-contaminated 3.2. Soret diffusion


oil samples.
Soret diffusion or the Soret effect refers to thermal diffusion,
which accounts for mass separation caused by the existence of a
3. Wax deposition mechanisms temperature gradient within the pipeline (Ekweribe et al., 2009).
Some researchers, such as Merino-Garcia et al. (2007), have classi-
The behaviour of waxy crudes is usually approximated by mod- fied its effect in wax deposition as negligible. However, expressing
elling them as Bingham-like fluids. Different mathematical models diffusion in terms of molecular and thermal diffusion allows for a
have been proposed ranging from a general one-dimensional mod- wax deposition model to more correctly account for thermal ef-
el of a waxy crude oil to models that describe crude oils depositing fects in diffusion (Banki et al., 2008). Thus, total mass flux would,
wax in closed flow loops. For example, Fusi (2003) and Fasano et al. ideally, have to be represented as a combination of Fick’s Law, in
(2004) delineate many models of differing complexity for the rep- terms of Dm and the concentration gradient, and transport by the
resentation of waxy crude oils. In order to fully model the flow of Soret effect, in terms of a thermo diffusion coefficient, DT, and
these crude oils, the mechanisms governing the deposition and re- the temperature gradient.
moval of solid wax must be incorporated into the model. Then
models can be developed, informed by a theoretical understanding 3.3. Brownian diffusion mechanism
of the mechanisms at play and the properties of the mixtures un-
der study. However, the question arises of which mechanisms This would occur when wax crystals that have precipitated out
are actually relevant. of the oil solution collide with excited oil molecules. The use of this
Investigations in this area have been ongoing for decades by mechanism in modelling deposition was also explored by Azevedo
researchers such as Hunt (1962); Burger et al. (1981), and Leiroz and Teixeira (2003). This diffusion mechanism can also be repre-
and Azevedo (2005). Azevedo and Teixeira (2003)did a critical re- sented by Fick’s Law as shown in equation.
view of wax deposition mechanisms, starting with wax deposition 
by molecular diffusion as described by Burger et al. (1981). In this dmB dC
¼ qd DB A ð2Þ
review it is acknowledged that, in most models of wax deposition, dt dr
molecular diffusion is treated as the dominant mechanism, and it is Here mB is the mass of wax deposited by Brownian motion, DB is the
also argued that experimental evidence suggests that gravity set- Brownian motion diffusion coefficient of the solid wax crystals and
tling and shear dispersion play no significant role in wax deposi- C is the concentration of solid wax out of solution.
tion. However, Azevedo and Teixeira point out that shear Azevedo and Teixeira (2003) acknowledge that many authors
dispersion may play a role in wax deposit removal, which would dismiss Brownian diffusion as a relevant mechanism for wax depo-
affect the rate at which wax accumulates. Other authors, such as sition. However, they conclude that there is not enough evidence
Solaimany Nazar et al. (2005b) and Correra et al. (2007), have to warrant this, citing an argument used by Majeed et al. (1990),
incorporated wax removal mechanisms involving shear forces which suggests that Brownian diffusion flux will be away from
(sloughing, ablation) into their wax deposition models. Other the wall, where the solid concentration would be highest. They dis-
mechanisms including thermo phoresis, the Saffman effect and miss this argument, because if the wax crystals are trapped in the
turbophoresis have also been considered in modelling wax deposi- immobile solid layer at the wall, the concentration of solid crystals
tion (Merino-Garcia et al., 2007). in the liquid at the wall is zero, or nearly zero, allowing for Brown-
ian diffusion toward the wall. The review concludes that Brownian
diffusion remains a possible contributing mechanism for wax
3.1. Molecular diffusion mechanism
deposition.
It is assumed that, for the flow of crude oil in the turbulent re-
gime, the turbulent diffusivities of momentum, chemical species 3.4. Gravity settling mechanism
and temperature will lead to a uniform distribution of velocity,
temperature and concentration profiles in a pipe cross-section. Azevedo and Teixeira (2003) classify gravity settling as insignif-
Therefore, the transport of wax will be controlled by the gradients icant in contributing to wax deposition, citing experimental evi-
prevailing at the laminar sub-layer close to the wall (Azevedo and dence from Burger et al. (1981), which showed that the settling
Teixeira, 2003). In a subsea pipeline in which the walls are cooled velocities of wax crystals under typical conditions do not contrib-
below the cloud point, there will be a radial temperature gradient ute significantly to deposition. This was further supported by
and wax crystallization will occur in cooler regions nearest to the experimental evidence from Burger et al., which demonstrated that
wall. Thus, solid wax crystals will exist in equilibrium with the li- deposition under horizontal and vertical flow is identical within
quid phase. Since wax solubility decreases with temperature, there the limits of experimental error.
will also be a concentration gradient established by the tempera-
ture gradient within the pipeline, with the cooler regions near 3.5. Shear dispersion mechanism
the wall having the lowest concentration of wax in the liquid
phase. This is what leads to the molecular diffusion of wax from Shear dispersion could contribute to wax deposition through
the bulk fluid to the walls of the pipeline. the lateral motion of particles immersed in a shear flow. Some
Azevedo and Teixeira (2003) suggested that the mass flux of the authors, such as Fusi (2003), include deposition in terms of a shear
wax be estimated by Fick’s Law as dispersion coefficient in the modelling of wax deposition. Also,
Fasano et al. (2004) claim that, based on the literature, for temper-
dmm dC atures much lower than the cloud point and for moderate heat
¼ qd Dm A ð1Þ fluxes the dominant process is shear dispersion, while for slightly
dt dr
higher temperatures the dominant process is molecular diffusion.
Here mm is the mass of deposited wax, qd is the density of the solid However, Azevedo and Teixeira (2003) claim that shear dispersion
wax, Dm is the diffusion coefficient of liquid wax in oil, A is the sur- does not contribute to deposition, because experimental evidence
face area over which deposition occurs, C is the concentration of shows no deposition of wax under conditions of zero heat flux,
wax in solution (volume fraction), and r is the radial coordinate. when it would only be possible if driven by a flow-induced
674 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

mechanism, such as shear dispersion. However, Azevedo and Paso (2005) also investigated the mechanical properties of waxy
Teixeira concede that shear forces can still contribute to the model fluids at constant cooling rates using controlled-stress rheo-
removal of wax deposits. Regardless of conflicting theories with metric measurements, applying an oscillatory upon the fluid sam-
regard to the role of shear dispersion in wax deposition, the impor- ples in order to characterize their mechanical properties during
tance of this mechanism in the overall accumulation and aging of gelation. The crystal structure in samples was also studied via
wax deposits cannot be ignored. microscopy and, furthermore, Paso applied an extension to an
established three-dimensional analytical percolation approxima-
tion to wax–oil gel systems. This allowed for the prediction of the-
3.6. Shear stripping mechanism
oretical gelation via the percolation threshold, the fractional
volume of the solid crystalline phase at which it forms a continu-
Removal of wax deposits by shear forces becomes especially
ous, domain-spanning path connected by crystal–crystal interac-
important under turbulent conditions when the rate of removal
tions. For this purpose, paraffin crystals were represented by
will be significantly higher compared to laminar flow. Therefore,
ellipsoidal geometries with spherical rotational volume of interac-
in order to accurately model wax deposition, especially for turbu-
tion. The primary and secondary ellipsoidal aspect ratios of the
lent flow, it is necessary to incorporate shear stripping effects into
crystals, a1 and a2, were related to the solid phase fraction at the
the model. Additionally, modelling wax removal by shear forces
percolation threshold, ug, by equation.
could help in the design of flow improver chemicals, as some of
these may act by causing the formation of softer gel structures that 1 1
/g ¼ hp ð3Þ
are more susceptible to removal by shear forces. Some researchers, a1 a2
such as Matzain (1999), have tried to represent this effect as an
Here hp = 0.295 represents the spherical percolation threshold.
empirical correlation for the reduction in rate of deposit formation
While this would give a prediction of the formation of a crystal per-
caused by shear forces.
colation network, it was noted that this will lead to gelation only if
the number density and strength of the crystal–crystal interactions
3.7. Nucleation and gelation kinetics are sufficient to impart solid-like properties to the fluid (Paso,
2005).
The crystallization of waxes is a kinetic process, the onset of Overall, Paso (2005) noted that the gel point of a waxy petro-
which can be described by classical homogeneous nucleation the- leum fluid is dependent on the morphologies and surface charac-
ory (Paso, 2005). While much work has been done to approach wax teristics of the randomly oriented paraffin crystals, and that
deposition as a thermodynamic problem, modelling based on the aspect ratios on the order of 100 allowed mechanical gels to form
kinetics of deposit formation has not been widely explored from these oils with paraffin content as low as 0.5%. Also, that
(Merino-Garcia et al., 2007). Paso (2005) sought to address the mono disperse crystals exhibited ordered surfaces and sharp edges,
insufficient understanding of the crystallization and gelation pro- providing minimal crystal–crystal contact and weak interactions,
cesses, as well as the assumption that paraffin precipitation kinet- while polydisperse n-alkane crystals exhibited nano-scale surface
ics does not limit deposition rates; an assumption that could lead roughness, which provides contact points for strong crystal–crystal
to the prediction of wax deposition in cases where a stable gel interactions, allowing for mechanical gelation at smaller wax con-
cannot form. He used model fluids consisting of n-paraffin compo- tents. Additionally, Paso concluded that percolation threshold
nents dissolved in petroleum mineral oils, and applied homoge- models provide accurate gel point predictions for physical gelation
nous nucleation and crystallization theory, along with differential systems that exhibit strong crystal–crystal interactions, while un-
scanning calorimetry to measure the onset of crystallization and der-predicting the solid fraction necessary to induce gelation in
the crystallization rate. weakly-interacting particle systems.
Paso (2005) compared experimental and equilibrium crystalli- Other recent studies that approached the subject of nucleation
zation rates to show that there were three regimes in the crystal- and gelation kinetics include those by Lopes-da-Silva and Coutinho
lization process at low cooling rates. The first is a nucleation lag (2007) and Ekweribe (2008). They analyzed gelation kinetics with
period starting at high-temperature conditions. The second is a the phenomenological Avrami model and noted an apparent
supersaturation growth period, driven by the supersaturation dependence of nucleation and crystal growth mechanisms and
established during the nucleation lag period as well as by decreas- rates on the degree of supercooling below the WAT at which crys-
ing solubility conditions, and during which the crystallization rate tallization is occurring. Lopes-da-Silva and Coutinho (2007) also
can spike well above the equilibrium crystallization rate. The third, noted an apparent predominance of heterogeneous nucleation
meanwhile, is an equilibrium growth period, which starts when and diffusion-controlled growth, especially at higher supercooling
the supersaturation ratio is diminished and the crystallization rate and/or higher oil complexity composition and molecular weight.
converges with the equilibrium predictions of the van’t Hoff rela- These results and those of Paso (2005) and further studies should
tion. One thing noted by Paso about these regimes was that the prove invaluable in the development of more robust wax deposi-
temperature span of the supersaturation growth regime was tion models, which take kinetic considerations into account. They
independent of the model fluid viscosity, providing evidence of can also be useful in determining mechanisms by which wax gela-
the absence of transport limitations in the crystallization rate. tion can be inhibited or wax deposits weakened by wax crystal
Through the application of the van’t Hoff solubility model with- modification.
in the framework of classical homogeneous nucleation theory, Paso
(2005)demonstrated that nucleation represents the primary 3.8. Deposition in two-phase flow
kinetic limitation associated with the crystallization of n-alkanes
in organic solution at low cooling rate conditions, with crystalliza- Analyzing and modelling liquid–liquid two-phase flow has pre-
tion rate limitations becoming significant at high cooling rates. He viously been explored by many researchers as well as present
also highlighted that the initial nucleation event is dependent upon authors (Raj et al., 2005; Chakrabarti et al., 2006, 2007). Deposition
the solubility behaviour of the highest fraction of n-alkane compo- in two-phase flow shows some characteristics similar to liquid–
nents in the fluid, and that the introduction of chain-length varia- liquid two-phase flow. Matzain et al. (2002) found that the
tions effects a reduction in the critical nucleus surface energy by thickness, hardness and profile of wax deposition in two-phase
co-crystallization of dissimilar chain-length paraffins. gas–oil flow show dependence on flow patterns. They used a closed
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 675

flow loop and the liquid displacement–level detection (LD–LD) entrapped dispersed water, as shown in Fig. 2. They observed a
technique, proposed by Chen et al. (1997), to measure wax depos- sharp increase in shear viscosity, yield stress and pour point for
its under different conditions. For horizontal flow, the thickness of waxy crude oil emulsions with above 25–30% volume of dispersed
deposits varied around the circumference of the pipe depending on water, as demonstrated in Fig. 3.
flow pattern, as shown in Fig. 1. It was similarly noted by de Oliveira et al. (2010) that these vis-
Matzain et al. (2002) account for these distributions by describ- cous emulsions can increase gel strength and hinder pipeline re-
ing how, in stratified flow, only the lower part of the wall will be in start by increasing the magnitude of the rheological properties of
contact with the oil phase, and the heat transfer rate will be high- the waxy crude oil gel. They attributed this change to the network
est at the bottom of the pipe and will decrease upward, resulting in developed by the aggregation of the waxy crystals and water. Paso
decreasing deposit thickness in a crescent shape. In the case of et al. (2009c) also noted drastic increases in fluid viscosities and
wavy stratified flow, the wavy gas–oil interface is cooled because shear thinning rheological behaviour due to the presence of emul-
of the waves, increasing heat transfer rate and, thus, deposit thick- sified water. These observations show the significance of consider-
ness along the interface. With intermittent flow, the passing of li- ing the effect of emulsified water on gelation, and Visintin et al.
quid slugs induces high shear force and stress along the bottom of (2008) note the importance of accounting for the impact of emul-
the test pipeline and shearing of wax deposits, resulting in thinner sified water during field development studies. Water fraction pro-
deposits at the bottom of the pipe. With annular flow, the wax duced by a well generally increases over its lifetime (Lockhart and
thickness is uniform around the circumference, as oil is uniformly Correra, 2005; Visintin et al., 2008). Thus it would be very useful to
in contact with the entire wall surface. account for the increasing impact of emulsified water on gelation
The results of Matzain et al. (2002) also showed changes in and gel rheology during continued operation.
hardness of the wax deposits for different flow patterns. Stratified
flow gave a soft deposit at the bottom of the pipe, with harder and
thicker deposits along the edge of the wavy gas–liquid interface. 5. Cloud point, pour point and gel point correlations
Intermittent flow resulted in a hard deposit, with increasing hard-
ness from the top to the bottom of the pipe. Lastly, annular flow re- Some authors have focused on developing correlations between
sulted in a very hard deposit, uniform across the circumference of measurable properties of crude oils, such as the pour point, and the
the pipe. Their results for vertical two-phase flow, on the other conditions under which disruptive wax deposition will occur.
hand, showed very uniform thickness distribution in the different Work such as this may help in predicting if and when fatal wax
flow regimes, with very hard deposits for annular flow, deposits deposition would occur in pipelines carrying particular crudes. Li
of medium to high hardness for intermittent flow, and hard depos- et al. (2005) cited the results of Holder and Winkler (1965) as indi-
its for bubbly flow with high superficial velocity. cating that 2 wt.% precipitated wax is sufficient to cause gelling of
virgin waxy crudes. Li et al. thus started with previously developed
correlations and tried to develop their own correlation linking the
4. Effect of emulsified water on gelation temperature at which this 2 wt.% precipitation would occur, Tc (2
wt%), and the pour point, Tpp, and gel point, Tgp, of various waxy
Crude oil emulsions, in particular, can pose significant flow crude oils., represented graphically by Figs. 4 and 5. These results
assurance risks and, with the increase in multiphase production and future research could be useful in both determining the ten-
in offshore environments, it has become important to evaluate dency for different waxy crudes to gel and harden at particular
the impact of emulsified water on crude oil gelation (Visintin temperatures, and in devising chemical means of inhibiting this
et al., 2008). The presence of water over a threshold value can pro- occurrence.
mote gel formation and viscous wax–oil gel emulsions. These
emulsions may be stabilized by the presence of polar compounds
such as asphaltenes and resins, and can have water cuts as high 6. Review of some existing wax deposition models
as 70% (de Oliveira et al., 2010). Paso et al. (2009c) attributed the
stability of waxy emulsions to the stabilizing effect of asphaltene Many different authors have proposed models for the flow of
particles on oil–water interfaces. They also suggested that, at waxy crude oils and the associated deposition of solid wax within
low-temperature conditions, molecular asphaltene adsorption pipelines, including Farina and Fasano (1997), and Fusi and Farina
onto precipitated wax crystals may increase the water wettability (2004). Additionally, there are commercial software codes
of the crystals, thus promoting adsorption at the oil–water developed to describe these processes, such as those compared
interface. by Bagatin et al. (2008). Fasano et al. (2004) reviewed various
Visintin et al. (2008) hypothesized that the solid paraffin stabi- mathematical models for the flow of waxy crude oils in laboratory
lizes the emulsion by being strongly adsorbed at the liquid–liquid experimental loops, in which the oils are assumed to behave like
interface forming Pickering emulsions. They suggested that, by non-Newtonian Bingham fluids, a common assumption for model-
means of the strong interaction between wax crystals and the drop ling these fluids. Torres and Turner (2005) approached the problem
surface, growth of the gel network involves the droplets them- by developing a method of straight lines for solving a Bingham
selves, forming a volume-spanning wax crystal network with problem for modelling the flow of waxy crude oils.

Stratified Stratified
Smooth Wavy Intermittent Annular

Fig. 1. Approximation of wax thickness distribution for various horizontal flow patterns (as described in Matzain et al., 2002).
676 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Fig. 2. Schematic representation of the gelation of waxy crude oil emulsions. Paraffin crystals that precipitate after a decrease of temperature below the WAT can adsorb on
droplet surface (A) or cover it (B), and stabilize the emulsion. Flocs of solid paraffin continuously grow on drops of water or between them (C). Dispersed water is entrapped
by a wax crystal network (D): the system spans the entire volume and the gelation is complete (Visintin et al., 2008).

Fig. 4. Tc (2 wt%) vs. ASTM pour point (Li et al., 2005).

concentration can be determined using the chain rule. This is a


problem that has been corrected in more recent deposition models
such as the one used in the Michigan Wax Predictor developed by
Fig. 3. Pour point of waxy crude oil emulsion with increasing water content
(Visintin et al., 2008). Hyun Su Lee.

6.1. Thermodynamic models


The earlier models presented here incorporate the wax deposi-
tion processes for pipelines containing waxy crude oils, and con- Many researchers have studied the thermodynamics of wax
sider cases where either molecular diffusion or shear dispersion deposition in hopes of creating a model that accurately describes
is considered the dominant mechanism involved in wax deposi- the process. In one example of earlier work, Lira-Galeana et al.
tion. However, one of the mistakes commonly introduced to wax (1996)developed a thermodynamic framework for calculating
deposition models is the assumption that the temperature and wax precipitation in petroleum mixtures as several distinct solid
concentration gradients are independent, and that, therefore, wax phases. Solaimany Nazar et al. (2005a) later developed a
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 677

and liquid phase respectively. For their model, there was an added
level of specificity for modelling particular n-alkane species. Differ-
ent correlations were used for the fusion enthalpies of n-alkanes
and for their enthalpies of solid–solid transition based on both car-
bon number and whether that number is odd or even. Similarly,
transition enthalpies were calculated for different components
based on chain lengths.
Table 2 shows a comparison of experimentally determined
WATs for three crude oils and those predicted by the model of
Wuhua and Zongchang (2006) and a similar model developed by
Leelavanichkul et al. (2004) and Fig. 6 compares the predictions
of the two models to experimental data for wax precipitation as
a function of temperature. The data indicates that refinement of
thermodynamic correlations, as performed by Wuhua and
Zongchang, can increase model accuracy in predicting precipita-
Fig. 5. Tc (2 wt%) vs. gel point (Li et al., 2005).
tion as a function of temperature.
Further studies, for example, by Edmonds et al. (2008), have
multi-solid phase thermodynamic model for predicting wax pre- also explored ways of representing the wax phase in order to more
cipitation in petroleum mixtures, by using the Peng–Robinson accurately model wax deposition. Edmonds et al. modelled the
equation of state to evaluate the phase behaviour of both liquid wax phase as a continuous distribution of n-alkane components,
and vapour phases. The model is solved for equilibrium, in which showing how this eliminated physically unrealistic artefacts found
the fugacity of each component is equal in every phase, using Eq. in the predictions of models that lumped n-alkanes into pseudo
(4), proposed by Prausnitz et al. (1986). components. Edmonds et al. carried out simulations with numbers
 s " ! ! of components approaching 100 and, in order to increase the com-
f DHfi T DHtri T putational speed, converted phase equilibrium and physical prop-
¼ exp 1   1 
fl i RT T fi RT T tri erty data into empirical expressions, fitted to the rigorous model.
Z Tf Z f # They also noted the importance of considering the deposit limiting
1 i 1 T i DC pi
þ DC pi dT þ dT ð4Þ mechanism of wax shearing in order for both their model and oth-
RT T R T T ers from the literature to more accurately agree with the limited
Here fis is the solid phase fugacity, DHri is the enthalpy of solid– field data available from actual pipelines.
solid transition between different solid phases, T fi is the tempera-
ture of fusion, Ttr is the transition temperature, Cp is the heat capac-
6.2. Hydrodynamic model
ity, and R is the ideal gas constant.
Table 1 shows a comparison of experimentally determined
Ramírez-Jaramillo et al. (2001) also developed a multi-solid
WATs for five synthetic paraffin systems and those predicted by
phase thermodynamic model for predicting wax deposition. In
the model of Solaimany Nazar et al. (2005a) and a UNIQUAC model
addition, Ramírez-Jaramillo et al. (2004) developed a multi-
developed by Coutinho (1998). The synthetic systems were each
component liquid-wax hydrodynamic model for simulating wax
composed of decane and a bimodal paraffin distribution. It should
deposition in pipelines, which treated molecular diffusion as the
be noted that with this and other models which use experimentally
dominant mechanism. Fig. 7a shows the computational domain
determined cloud points to validate the model, there is a limit to
how accurately cloud points can be measured which is highly
dependent on the particular oil mixture, as discussed by Coutinho Table 2
and Daridon (2005) and Hammami et al. (2003). Therefore, agree- Experimental WAT data and model predictions for crude oils (Wuhua and Zongchang,
2006).
ment with experimental data may not prove definitively the accu-
racy of a model, especially as far as its applicability to a wide range Sample Experimental Leelavanichkul Deviation Present Deviation
of wax–oil mixtures. results model model
Wuhua and Zongchang (2006) also developed a more recent Crude 298.2 K 298.8 K 0.4 K 301.3 K 3.1 K
thermodynaamic model, based on the equality of fugacities at Oil A
Crude 295.2 K 293.4 K 1.8 K 295.4 K 0.2 K
equilibrium, which estimates solid precipitation as a function of
Oil B
temperature and composition. For this study, Eq. (5) was used for Crude 294.2 K 296.0 K 1.8 K 297.8 K 3.6 K
the condition of equal fugacities in the solid and liquid phases. Oil C
Z !
xSi cL f L P
V Li  V Si
L
¼ iS iS exp dP ð5Þ
xi ci fi 0 RT

Here x is the mole fraction, c is the activity coefficient, V is vol-


ume, P is pressure and the S and L superscripts indicate the solid

Table 1
Comparison of the WAT between experimental data, UNIQUAC model and Solaimany
Nazar et al. model (Solaimany Nazar et al., 2005a).

Bim 0 Bim 3 Bim 5 Bim 9 Bim 13


WAT (K) 308.75 309.65 310.37 311.33 312.81
UNIQUAC 307.05 307.55 308.47 309.63 311.41
This model 308.45 309.05 309.55 310.7 312.75 Fig. 6. Wax precipitation as a function of temperature for crude oil A (Wuhua and
Zongchang, 2006).
678 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Fig. 7. (a) Computational domain for a model pipe. (b) Sections of a model pipe with concentric layers (Ramírez-Jaramillo et al., 2004).

used by Ramírez-Jaramillo et al. (2004), consisting of a model pipe Ramírez-Jaramillo et al. (2004) used mass, momentum and
of length, L, and radius, r, along which a mixture of hydrocarbons energy balances, shown in Eqs. (6)–(8) and assumed mixture
flows. The pipe was divided to form a computational mesh, with incompressibility and quasi-steady state for all rate processes
boundary conditions applied at the ends and along the exterior concerning mass, momentum and energy.
surface of the pipe, and finite differences were used in the solution
of differential equations. @ qm
Ramírez-Jaramillo et al. (2004) modelled the fluid as consisting þ r  qm m ¼ 0 ð6Þ
@t
of n hydrocarbon components in thermodynamic equilibrium, with
 
mole fractions, in both the liquid and solid phases, that are func- @m
tions of pressure and temperature. They considered the wax depo- qm þ m  rm ¼ rP þ r  s þ qm g ð7Þ
@t
sition rate to depend on oil composition, oil temperature, external
temperature around the pipe, flow conditions, pipeline size and  
@T
pressure. The model assumed wax deposition by molecular diffu- qm C v þ m  rT ¼ kr2 T ð8Þ
@t
sion and removal by shear forces, which would be especially signif-
icant at high Reynolds numbers [( quDh /l), where q = density,
u = velocity, l = dynamic viscosity, Dh = hydraulic diameter]. In Here P, s and g are the pressure, stress tensor and gravitational
addition, the model included aging by the diffusion of wax into constant; Cm and k are the heat capacity and thermal conductivity
and within the gel-like deposit, which is discussed later in this (which is assumed constant), respectively; and m is the average
paper. The mass flux was calculated for all components in the macroscopic velocity of the mixture. They expressed the total
system and summed to give the total flux. amount of deposited wax, M(t,z), in terms of the deposited mass
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 679

of each component due to molecular diffusion, MMDi(t,L), the mass the cloud point, while pumping the wax–oil mixture through the
removed by shear forces, MSR(t,L), and the mass of wax molecules flow loop. The flow loop consisted of a 5/8 in. OD steel tubing test
diffusing into the gel deposit, MGD(t,L), as shown in equation. section, which was cooled by a heat exchange jacket, and an iden-
tical but non-cooled reference section. Pressure taps connected to
X
n
Mðt; zÞ ¼ M MDi ðt; LÞ  M SR ðt; LÞ  MGD ðt; LÞ ð9Þ pressure transducers were used to measure the increase in differ-
i¼1 ential pressure during operation in order to determine the thick-
ness of the deposit within the test section. The bulk fluid inlet
Ramírez-Jaramillo et al. (2004) solved for the total deposition
temperature, tb, and wall temperature, ta, were also monitored.
rate, @M/@t. The output of their model included solid fractions, den-
Singh et al. (2000) determined that a counter diffusion phenom-
sity, viscosity, radial mass flux and deposited mass calculations.
enon, in which wax molecules diffuse into the gel deposit and oil
The model showed reasonable agreement with previously devel-
molecules diffuse out of the deposit, is responsible for the aging
oped models and experimental data for a binary mixture reported
of the deposit. They furthermore determined that the rate of aging
by Cordoba and Schall (2001), as shown in Fig. 8 Ramírez-Jaramillo
is dependent on oil flow rate as well as the pipeline wall tempera-
et al. found that the Peclet number and Reynolds number parame-
ture. In their experimental setup with oil in a closed flow loop with
ters had a significant impact on the amount of wax deposited.
cooled walls, there was a rapid decrease in internal radius mea-
sured over the first day followed, which then plateaued. Similarly,
7. Wax aging models the increase in the measured weight fraction of wax slowed after a
rapid change in the first day. The wax content (determined using
7.1. Counter diffusion high-temperature gas chromatography, HTGC) of the gel deposit
also changes over time, with the proportion of lighter components
Researchers have also explored the properties of wax crystals decreasing after the first day, while the proportion of heavier com-
and wax deposits formed from crude oils. Nautiyal et al. (2008) ponents increases. The data recorded by Singh et al., showed that
studied the crystal structure of n-alkane paraffins crystallized from the wax content of the deposit continued to increase even after
crude oil. Other studies have specifically looked at the way wax the thickness stabilized, and that waxes of chain length higher than
deposits change after the initial formation. This is important be- 29 diffused into the deposit while the ones with lengths less than
cause, in addition to understanding the mechanisms involved in 29 diffused out. 29 is the critical carbon number, CCN, for the given
the deposition of wax in pipelines, in order to fully model the flow operation conditions; a value that could be useful in determining
of crude oil and accumulation of wax, it is vital to understand the what inhibitors to use in a particular well or pipeline, based on
mechanisms that govern the aging of wax deposits. These deposits whether or not they can inhibit crystallization of waxes above
are not simply static and unchanging. Rather, after a layer of wax the CCN (Paso and Fogler, 2003).
has formed along a pipeline wall, its composition gradually Singh et al. (2000) were able to develop a mathematical model
changes. The crystalline wax deposit actually behaves like a porous to describe the wax deposition process in a laboratory flow loop by
medium with oil trapped within its three-dimensional network solving numerically a coupled system of differential and algebraic
(Singh et al., 2000, 2001a). The wax content of this deposited gel equations of heat and mass transfer inside and outside the gel de-
can therefore increase with time by diffusion. As this happens, posit. Eq. (10) shows the mass balance they used to relate the rate
hardness, melting point and heat of fusion of the deposit can of change of wax in the gel deposit to the radial convective flux of
change, which could affect decisions about the appropriate method wax molecules from the bulk of the fluid–gel interface.
of wax removal to employ in a pipeline.
d
Singh et al. (2000) studied this phenomenon by use of food ½pðR2  r 2i ÞF w ðtÞLqgel  ¼ 2pr i Lk1 ½C wb  C ws ðT i Þ ð10Þ
grade wax dissolved in a mineral oil–kerosene mixture, which dt
was pumped through a closed flow loop setup. Their experimental Here R is the original internal radius of the pipe, ri is the internal
procedure involved heating a wax–oil mixture to 30–35 °C in a radius during deposition (average radius available for flow of oil),
stirred tank and maintaining the temperature of this vessel above F w is the weight fraction of solid wax in the oil, L is the length of
pipe, qgel is the density of the gel deposit (considered constant),
k1 is the mass transfer coefficient, Cwb is the bulk concentration
of wax, Cws is the solubility of the wax in the oil solvent derived
in terms of Ti, and Ti is the interfacial temperature, which was ob-
tained from the energy balance shown in equation,
2pke ðT i  T a Þ
2pr i hi ðT b  T i Þ ¼  2pri k1 ½C wb  C ws ðT i ÞDHf ð11Þ
lnðR=r i Þ
where hi is the interface heat transfer coefficient, ke is the effective
thermal conductivity of the gel, and DHf is the heat of solidification
of the wax. The heat and mass transfer coefficients were obtained
using Hausen, Seider and Tate correlations.
Eq. (12) shows the deposit growth equation derived by Singh et
al. (2000), by relating the rate of addition of wax to the gel deposit
in the flow loop to the radial convective flux of wax molecules from
the bulk to the fluid–gel interface and the diffusive flux into the gel
at the gel interface.
 
dr i dC w 
2pr i F w ðtÞqgel ¼ 2pr i k1 ½C wb  C ws ðT i Þ  2pri De ð12Þ
dt dr i

Fig. 8. Dimensionless wax thickness distribution vs. time. Comparison of model


Here De is the effective diffusivity of wax inside the gel deposit.
predictions with experimental data for the 30:70 (cyclo C6C19:C8) ratio (Ramírez- Coupled differential equations from Eqs. (10) and (12) were solved
Jaramillo et al., 2004). by Singh et al. throughout the length of the pipe at each time
680 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Kk
r¼ ð13Þ
b cos h
Coutinho et al. (2003) noted an increase in crystal size observed
by CPM at a temperature in the neighbourhood of the pour point.
They reported an increase from 6.4% to 15.3%, over 110 h, for the
fraction of a CPM image occupied by crystals. Furthermore, they
obtained Differential Scanning Calorimetry (DSC) thermograms
under the same conditions, which did not show detectable heat
effects associated to this change in the crystal size, as seen in
Fig. 9. Kinetic growth of crystals for oil sample from X-ray diffraction analysis at Fig. 10. They noted that this can only occur when the heat of crys-
10 °C (Coutinho et al., 2003). tallization released is used by the melting of an equivalent mass of
crystals. This indicates that wax deposits in crudes suffer recrystal-
lization. Coutinho et al., thus, conclude that Ostwald Ripening is
instant using Runge–Kutta algorithms, along with equations for
also a mechanism responsible for the aging of wax deposits.
dTi/dr. This system of equations was used to obtain the trajectories
of thickness and wax content at each location in the pipe. This
model showed excellent agreement with experimental data. How- 8. Correct analogies for correlated heat and mass transfer in
ever, Lee (2008) has shown that the mass-heat transfer correla- turbulent flow
tions used by Singh et al. incorrectly assume independent heat
and mass transfer, and were successfully applied only because of Many existing wax deposition models assume that heat and
a high degree of supersaturation in the laminar boundary layer. mass transfer can be related by the chain rule, which assumes that
Other researchers, such as Hernandez et al. (2004), modelling the system is at thermodynamic equilibrium (which may not be
wax deposition in pipelines has begun incorporating wax aging true), or use mass-heat transfer analogies, such as the Chilton–
into their models. Additionally, Singh et al. (2001b) were able to Colburn analogy, which are valid only when the temperature and
develop a thermodynamic model to predict both cloud point tem- concentration fields are independent. Venkatesan and Fogler
peratures and CCNs of wax–oil mixtures, where CCN is a function (2004) noted that such heat-mass transfer analogies are not appli-
of the mixture composition as well as the wall temperature. This cable for predicting the mass transfer rates in turbulent flows,
model also showed good agreement with experimental data, pre- where the concentration field is correlated to the temperature field
dicting the cloud points and the CCNs of model oils with good and the concentration boundary layer and temperature boundary
accuracy. layer thicknesses are not independent. They showed that use of
the Colburn analogy in this case would result in a significant
7.2. Ostwald ripening over-prediction of wax deposition. They also proposed a method
for estimating the convective mass transfer rate using the Nusselt
It must be noted that the diffusion mechanism used by Singh number and the experimentally obtained solubility curve. How-
et al. (2000, 2001a,b) is not the only possible mechanism for ever, this method would only be valid for thermodynamic equilib-
explaining the aging process. In fact Continuo et al. (2003) found rium in the mass transfer boundary layer, when precipitation
that aging of wax deposits takes place even for samples kept under kinetics are not limiting.
isothermal conditions. The diffusion mechanism for aging cannot For the development of more rigorous and accurate models, it
account for this as it is driven by temperature-composition gradi- has been necessary for researchers to explore the correct relation-
ents. Coutinho et al. reported broadening of peaks on X-ray diffrac- ship between heat and mass transfer. Lee (2008) investigated the
tion and Cross Polar Microscopy (CPM) images showing an increase combined heat and mass transfer phenomenon under laminar
in the crystallite’s size. Fig. 9 shows an example of their results and turbulent flow conditions using the finite difference method.
from X-ray diffraction analysis, for which the crystallite size, r, is He developed a model based on that of Singh et al. (2000), which
related to a shape factor K, and the measured peak position, h, could be applied for any precipitation kinetics. For turbulent flow,
and breadth, b, by equation. Lee showed that the solubility method proposed by Venkatesan

Fig. 10. Thermogram for oil C (thick line). The isothermal region above 5000 s shows that there are no detectable heat effects related to the aging of the wax (Coutinho et al.,
2003).
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 681

and Fogler (2004) under-predicts deposition by assuming that the In his model, after calculating the Sherwood and Nusselt num-
concentration profile in the mass transfer boundary layer follows bers, Lee (2008)could then solve the growth and aging governing
the thermodynamic equilibrium limit between temperature and equations from Singh et al. (2000)’s model to solve for deposit
concentration at every point. This was contrasted with the over- thickness and wax fraction at each time step in his computational
prediction of the Chilton–Colburn analogy, which gave maximum procedure. Lee (2008)showed that there was excellent agreement
supersaturation. The comparison showed that these two ap- between the results of his model and lab-scale laminar flow loop
proaches constitute the limiting cases for deposition, and that experimental data. There was also good agreement with turbulent
the actual concentration profile, which is dependent on the precip- lab-scale results, though there was significant discrepancy for early
itation kinetics, falls between those calculated by the two methods. times at higher volumetric flow rates, possibly due to sloughing.
Instead of using those two limiting cases, Lee (2008) employed The results of the computational model also closely matched
a computational approach for calculating the Nusselt numbers large-scale flow loop data. The results obtained by Lee (2008) show
[(hL/kf), where L = characteristic length, kf = thermal conductivity that this model is applicable for varying precipitation kinetics, and
of the fluid, h = convective heat transfer coefficient] and Sherwood provides a robust and rigorous way of predicting wax deposition
numbers [(KL/D) where L is a characteristic length, D is mass diffu- under a range of turbulent conditions.
sivity, K is the mass transfer coefficient] according to following
equations.
9. Inhibition of wax deposition

ð2ri Þ@T 
@r rr ð2r i Þhi
Nu ¼ i
¼ ð14Þ The most effective way of dealing with the problem of wax
Tb  Ti k
deposition in crude oil pipelines would be to prevent it from occur-
 ring in the first place. Thus, researchers have investigated different
ð2r i Þ@C  methods of inhibiting the deposition process. These include the
@r rr ð2r i ÞkM
Sh ¼ i
¼ ð15Þ heat insulation of subsea pipelines to actually inhibit precipitation
Cb  Ci Dw0
by keeping pipeline temperatures as high as possible (Quenelle and
The temperature and concentration gradients at the fluid-de- Gunaltun, 1987), the internal coating of pipelines with plastics
posit interface, needed for these calculations, were obtained by (Patton, 1970; Bummer, 1971), and also methods of preventing
solving mass and energy balance equations, as shown in following wax deposition on pipeline walls, such as the use of chemical
equations. inhibitors, which will be discussed in more detail in this paper.
 
@C 1 @ @C
mz ¼ rDwo  kr ðC  C ws Þ ð16Þ
9.1. Chemical inhibitors
@z r @r @r

  Many researchers have studied the efficacy of different inhibi-


@T 1 @ @T
mz ¼ r aT  bðC  C ws Þ ð17Þ tors of wax deposition and the mechanisms by which they inhibit
@z r @r @r
this deposition, including Jorda (1966), Mendell and Jessen (1970),
Here vz is the axial velocity, Dwo is the molecular diffusivity of Fulford (1975), Addison (1984), Newberry and Barker (1985),
wax in oil, kr is the thermal conductivity, aT is the thermal diffusiv- Fielder and Johnson (1986), Singhal et al. (1991), Jang et al.
ity, and the precipitation term b(C–Cws) is considered negligible. (2007), and Tinsley et al. (2007). The efficacy of commercially
Lee first did this for laminar flow. Using a discretized form of the available inhibitors tends to be limited, and has to be evaluated
mass-heat transfer equation along with their appropriate bound- on a case-by-case basis. Wang et al. (2003), for instance, found,
ary conditions, Lee wrote the governing equations in matrix form. when testing some wax inhibitors, that the inhibitors they had
Then by inverting these matrices to give the radial temperature studied reduced the total amount of deposition, but had only lim-
and concentration profiles, and numerically marching from the in- ited success in suppressing the deposition of the high molecular
let of the tube to the exit he could obtain the complete set of tem- weight paraffin components (C35 and above). This resulted in hard-
perature and concentration profiles with respect to the radial and er wax deposits than in the absence of an inhibitor. They also found
axial position. that inhibitors most able to depress the WAT were more likely to
From this, Lee (2008) showed how the Sherwood number be superior products for decreasing total wax deposition, and that
profile as a function of axial distance would change for different the addition of the corrosion inhibitor, oleic imidazaline (OI), sig-
precipitation rate constants. This showed that if there was no pre- nificantly increased the efficacy of deposition inhibition. Fig. 11
cipitation in the boundary layer, the heat and mass transfer rates shows some of their results, where PIE is the paraffin inhibition
become independent of each other, resulting in a supersaturation efficiency, the amount of wax deposited with inhibitor as a wt.%
curve. However, as the precipitation rate constant increases the of amount deposited without it.
Sherwood number is decreased, because wax molecules would Bello et al. (2006) also studied the efficacy of commercial wax
not reach the oil–deposit interface, and would instead flow down inhibitors, particularly on Nigerian crude oils. They found that
to exit as solid particles. the use of a trichloroethylene–xylene, TEX, binary system as an
To obtain the Sherwood and Nusselt numbers under turbulent additive was actually more effective and economically feasible
conditions, Lee (2008) used the same procedure with governing than the use of commercial inhibitors. Other researchers have
equations modified for turbulent flow to include the turbulent noted the need to tailor inhibitor treatments to particular crudes
axial velocity profile and the thermal and mass transfer eddy diffu- in order to maximize efficacy. Manka and Ziegler (2001), for in-
sivities. The wax concentration profiles in the turbulent boundary stance, found that additives work best when matched to the paraf-
layer obtained this way showed that heat and mass transfer fin distribution in the crude oil being treated. Similarly, Carmen
become independent as the precipitation rate constant approaches García (2001) noted a strong relationship between a specific paraf-
zero, resulting in the Chilton–Colburn analogy-derived concentra- fin inhibitor’s efficiency and the crude oil composition, which
tion profile. Conversely, as the precipitation rate constant would require case-by-case consideration for selecting inhibitors
increases, precipitation in the boundary layer increases, with con- for use in the field.
centration approaching the solubility limit for thermodynamic Additionally, there is the consideration of the environmental
equilibrium. conditions under which a wax inhibitor is to be used, since, for
682 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Fig. 11. Effect of wax inhibitors (100 ppm) and oleic imidazoline, OI, (200 ppm) on paraffin deposition from a mixture of paraffin wax in C10 solution (Wang et al., 2003).

operations at particularly low temperatures, the inhibitor formula- polymers with long alkyl groups, such as alkyl phenol–formalde-
tion must be winterized to allow effective delivery under those hyde, which are not as effective as comb polymers when acting
conditions (Manka et al., 1999; Jennings and Breitigam, 2009). on their own as flow improvers (Kelland, 2009).
Also, while work continues towards developing new, more effec-
tive wax inhibitors, it remains the case that inhibitors typically 9.2.1. Ethylene copolymers
do not provide 100% inhibition, and so are used in conjunction with This group includes ethylene/small alkene copolymers, ethyl-
remediation methods such as pigging (Jennings and Breitigam, ene/vinyl acetate (EVA) copolymers, and ethylene/acrylonitrile
2009; Kelland, 2009). copolymers (Kelland, 2009). More specifically, examples of these
polymers used in wax inhibition studies include poly (ethylene-
9.2. Types of chemical inhibitors b-propylene) and poly(ethylene butene) polymers (Tinsley et al.,
2007; Kelland, 2009). Random, low molecular weight EVA copoly-
There are different mechanisms by which chemical inhibitors mers, illustrated in Fig. 12, are widely used and investigated as wax
can prevent wax deposition or gelling in pipelines. They can lower inhibitors (Kelland, 2009). The effectiveness of the EVA copolymer
the WAT or pour point or can modify the wax crystals so as to pre- as an inhibitor is influenced greatly by the percentage of vinyl ace-
vent their agglomeration and deposition (Kelland, 2009). The tate in the copolymer. The, more polar, vinyl acetate content aids
chemicals that modify the WAT are usually referred to as wax solubility and lowers crystallinity and so is necessary for the
inhibitors or wax crystal modifiers, while those that affect the pour depression of the WAT, whereas the polyethylene content is neces-
point are known as pour point depressants (PPDs) or flow improv- sary to allow for co-crystallization with structurally similar wax,
ers; although there is a great deal of overlap in terms of the chem- but, on its own, has little effect on crystallization (Kelland, 2009).
istry and mechanisms of these two classes (Kelland, 2009). Some
detergents or dispersants that act as wax inhibitors, such as poly- 9.2.2. Comb polymers
esters and amine ethoxylates, may act partly by modifying the sur- Comb-shaped polymers, illustrated in Fig. 13, have been studied
face of the pipe wall, rather than just the wax crystals, to prevent extensively as wax inhibitors by researchers such as Duffy and
adhesion (Pedersen and Rønningsen, 2003), and many effective Rodger (2002), Duffy et al. (2004), Jang et al. (2007), and Soni et
wax inhibitors create weaker deposits that are more easily re- al. (2008). They are usually made from (meth)acrylic acid or maleic
moved by shear forces (Manka et al., 1999; Kelland, 2009). The anhydride monomers, or both, and generally provide improved
main types of wax inhibitors and PPDs include ethylene polymers wax inhibition compared to the ethylene copolymers (Kelland,
and copolymers, comb polymers and assorted other branched 2009). One proposed mechanism for their action as PPDs is that

Fig. 12. Ethylene/vinyl acetate (left) and ethylene/acrylonitrile copolymers (right) (Kelland, 2009).
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 683

Fig. 13. Traditionally depicted structure of a comb polymer (left). X is a spacer group. The structure looking down the helical backbone (right) (Kelland, 2009).

comb polymers reduce the ability of wax crystals to agglomerate layer from which wax crystals are easily sheared off, or by adsorb-
into a gel structure by introducing defects or repulsive forces (Jang ing onto the wax crystals and reducing their tendency to stick
et al., 2007; Soni et al., 2008; Kelland, 2009). As illustrated by Figs. together (Kelland, 2009). Some researchers have worked on devel-
14 and 15, they can accomplish this by providing nucleating sites oping their own dispersant formulations. Groffe et al. (2001), for
for wax crystals on their paraffin-like pendant chains while a polar instance, developed their own inhibitor that shows wax dispersant
backbone impedes the formation of an interlocking wax network behaviour and anti-sticking properties. They suggest that this
(Soni et al., 2008). chemical, referred to as P5, interferes with the wax crystal growth
In selecting the most effective comb polymers for use with a mechanism by preventing the formation of a three-dimensional
particular crude oil, researchers have found that the length of the network, and thus reduces the pour point and improves the flow
side chains plays an important role. For example, Manka and characteristics of crude oils. Fig. 16 shows the effectiveness of P5
Ziegler (2001) found that matching the average pendant chain in preventing the adherence of wax from crude oils to a steel
length of comb polymer PPDs with the paraffin distribution of a surface.
crude oil provided the greatest pour point depression. Also, Jang Typical, low-cost wax dispersants include alkyl sulfonates, alkyl
et al. (2007) obtained results which suggested that using comb aryl sulfonates, fatty amine ethoxylates and other alkoxylated
structures with side arms of such length as to interact favourably products, but these dispersants have shown limited effectiveness
with the fraction of oil most likely to crystallize into the hard in the field when not blended with polymeric wax inhibitors (Kel-
wax phase provided the best wax inhibition. This creates a prob- land, 2009). Dispersants, however, have been used successfully to
lem for especially long-chained waxes, for which it would be diffi- support the functions of polymeric flow improvers because of their
cult to introduce a comb polymer (or ethylene copolymer) of ability to hinder wax settling and deposition (Al-Sabagh et al.,
sufficient length to provide efficient inhibition, and also makes it 2007).
important to have a range of comb polymers available for treat-
ment of different crudes (Kelland, 2009). 9.2.4. Polar crude fractions
It has been found that polar extracts from crude and distillate
9.2.3. Wax dispersants oils, which can be extracted using super critical gases such as car-
These are surfactants that adsorb onto pipe surfaces and reduce bon dioxide or ethylene, and which contain asphaltenes, resins and
the adhesion of waxes to those surfaces, possibly by changing the aromatics, can be a potential source of low-cost flow improvers
wettability of the pipe surface to water-wet, or by creating a weak (Kelland, 2009). Venkatesan et al. (2003) studied the effects of

Fig. 14. Characteristic structure of a comb polymer PPD (Soni et al., 2008).
684 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Fig. 15. Prevention of interlocking of wax crystals by polymer additives by (a) providing nucleating sites to asphaltene as well as wax molecules; (b) polar parts hinder the
co-crystallization of both wax as well as asphaltenes (Soni et al., 2008).

Fig. 16. Effect of 500 ppm of P5 on wax adherence to a steel surface exposed to three different crude oils (Groffe et al., 2001).

asphaltenes on the formation of paraffin gels in crude oil. They addition resulted in macroscopic phase separation of the mixture,
found that the addition of asphaltenes depressed the gelation tem- attributable to gravity settling.
perature of model wax–oil mixtures, as summarized in Figs. 17 and Kriz and Andersen (2005) also studied the effect of asphaltenes
18; although they found that beyond certain thresholds, further on wax crystallization in crude oils. They found that this effect
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 685

certain degree. These are all properties associated with PPDs and
proposed mechanisms for their action, which include co-precipi-
tating with waxes and hindering crystal network growth or coating
wax crystals to prevent agglomeration. Thus, as also suggested by
the observations of Kriz and Andersen (2005), it stands to reason
that asphaltenes affect wax precipitation by the same mechanisms
as other flow improvers such as comb polymers.

9.2.5. Short-chain alkanes


Senra et al. (2008) analysed how n-alkanes impact the crystalli-
zation of one another, and Senra et al. (2009) studied the gelation
characteristics of long-chained n-alkanes in a short-chained
n-alkane solvent, looking at the inhibition of gel formation caused
by the addition of other crystallizable n-alkanes to long-chained
n-alkanes, which are the primary component of wax deposits. As
is the case with polymeric inhibitors, the results obtained by Senra
Fig. 17. Gelation temperature depression (cooling rate of 1 °C min1) for a food- et al. (2009) indicate that the ability of a particular short-chained
grade paraffin wax (Wax 1) and a laboratory-grade paraffin wax (Wax 2) by n-alkane to inhibit gel formation by a longer-chained one depends
addition of asphaltene (Venkatesan et al., 2003).
on the particular pairing. The trend of this inhibition was found
to depend on the extent of differences in size and solubility char-
acteristics between the long-chained n-alkane and the added
shorter-chained one as demonstrated by the results in Figs. 19
and 20.
Senra et al. (2009) found that, for a given wax percent of a long-
chained n-alkane, polydispersity and co-crystallization weaken the
gel formed in spite of the fact that more crystallizable wax is pres-
ent in solution. In cases where co-crystallization was possible, such
as in a C36/C32 system, they witnessed a noticeable decrease in
pour point and gelation temperature with the addition of small
amounts of the shorter n-alkane. This, they accounted for by the
defects in the crystal structure that would be required to accom-
modate the C32 crystals that co-crystallize with the C36. This would
make the formation of large crystals and a volume-spanning
network gel more difficult, in the same way that the inclusion of
polymeric flow improvers into wax crystal structures inhibits
aggregation and gel formation. The addition of increasing concen-
trations of the shorter-chained co-crystallizing n-alkane, however,
resulted in a minimum pour and gel point followed by an increase.
Fig. 18. Depression in yield stress of Wax 1 system (at temperature, Tys, below the
This was accounted for by a limit to how much the addition of the
gelation temperature) upon asphaltene addition (Venkatesan et al., 2003).
shorter-chained n-alkane can decrease crystal size, beyond which
further addition only adds more material to form wax crystals.
On the other hand, with n-alkanes of similar size which did not
depends strongly on the degree of asphaltene dispersion or floccu-
co-crystallize, such as in a C36/C28 system, Senra et al. (2009) saw a
lation more than on the asphaltene type or origin. They reasoned
very different trend. In this case, low concentrations of the shorter-
that the asphaltenes, when well dispersed at very low concentra-
chained n-alkane had no effect on the pour and gel points, until a
tions, are easily accessible for any kind of interaction with the par-
affins and can be fully incorporated into the wax structure. They
noted a delay in crystallization, which indicated that building the
asphaltene molecules into this structure would require a higher
driving force because of asphaltene–paraffin spatial interference.
This would suggest that the asphaltenes are acting by some of
the same mechanisms proposed for inhibition by polymeric
inhibitors.
In agreement with the results of Venkatesan et al. (2003), Kriz
and Andersen (2005) also saw a depression in yield stress and
WAT, which they accounted for by suggesting that asphaltene mol-
ecules flocculate together when over a critical concentration, with
possible co-precipitation with waxes, resulting in an unorganized
asphaltene–paraffin composite rather than a proper wax network.
They note, though, the need for further understanding of the way
asphaltenes and waxes interact during wax crystallization, and an-
other study by Yang and Kilpatrick (2005) indicated that asphalt-
enes and waxes do not co-precipitate in solid organic deposits.
In accounting for the observed flow improver properties of
asphaltenes, Venkatesan et al. (2003) noted that asphaltenes have Fig. 19. Effect of varying the wax percent of C28 and C32 on the pour points and gel
polar groups as well as alkane chains and are soluble in oil up to a points of 4% C36 solutions in dodecane (Senra et al., 2009).
686 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

C32/C24 system, where co-crystallization does not occur, the C24


was seemingly too small to influence the crystal structure and
impact C36 gelation, and so simply acted like a solvent. These
results show that an understanding of how oil composition affects
wax–oil gel formation can help significantly in implementing
inhibition measures.

9.3. Surfaces that prevent wax deposition

There is an obvious appeal to developing wax-repellent surfaces


for use in oil pipelines as this would limit or eliminate the need for
wax inhibition and removal measures to maintain normal opera-
tion. With a proper understanding of the mechanisms by which
Fig. 20. Effect of varying the wax percent of C28 and C30 on the pour points and gel waxes adhere to oil pipeline walls it would be possible to create
points of 4% C32solutions in dodecane (Senra et al., 2009).
pipelines in which the nature of the walls makes adhesion unfa-
vourable. Paso et al. (2009a) performed a comprehensive review
concentration at which a sharp decrease was witnessed in both of the use of non-stick and anti-adhesive coatings for inhibiting so-
followed by a gradual increase. This was accounted for by the fact lid–liquid deposition phenomena, including the use of metal sur-
that, at low concentrations, the more soluble shorter-chained n- face treatments and synthesized polymers. The classes of
alkane will not crystallize out and will not be present in high materials that they found promising included fluoro-siloxanes, flu-
enough concentration to disrupt the crystallization of the longer- oro-urethanes, oxazolane-based polymers and hybrid diamond-
chained n-alkane, so will have no effect. Then, at a high enough like carbon and polymer coatings.
concentration, the association of the shorter-chained n-alkane Fig. 21 shows some of the reported surface free energies of sur-
molecules with the longer-chained n-alkane crystals would disrupt faces for paraffin control investigated by Paso et al. (2009a), which
gel formation. Then gelation will occur as the more soluble shorter- gives an indication of the ability for waxes to interact with those
chained alkane is added in high enough concentration to crystallize surfaces, and thus the potential of these surfaces for preventing
sufficiently to form a gel. Senra et al. supported this analysis with wax deposition. Further study of the mechanisms involved in
the results of cross-polarized microscopy experiments. wax adhesion, hopefully, will result in even more effective surface
Furthermore, for a C32/C30 system, Senra et al. (2009) noted that, treatments in the future.
due to the very similar chain length and solubility characteristics,
there was only a slight initial decrease in pour point due to the for- 9.4. Cold flow
mation of co-crystals, which would have relatively few vulnerable
points since the two n-alkanes are so similar. Beyond that, the C32/ Heating or insulation of subsea pipelines can be used to try to
C30 system behaved much like a monodisperse system. Also, for a prevent cooling of the pipeline wall below the WAT. However, a

Fig. 21. Surface energy reduction possible with novel surface technologies (Paso et al., 2009a).
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 687

very different method of inhibiting wax deposition on pipeline compressed ahead of the pig. In such an event the pipeline could
walls, discussed by Merino-Garcia and Correra (2008), is the use be lost. The use of bypass pigs tries to address this problem. When
of Cold Flow technology. This approach suggests that it might be the differential pressure across such a pig becomes too high, be-
possible to prevent deposition on pipeline walls by reducing the cause of the accumulation of solid wax and debris ahead of it,
bulk temperature within the pipeline to be equal to the tempera- the bypass pig allows liquid to flow through it and disperse the
ture of the sea water around it, thus eliminating the temperature accumulated solid ahead. However, there is always the danger that
gradient. This would allow for the waxes to be transported as a so- if pigging has to be temporarily suspended due to mechanical fail-
lid dispersion within the bulk fluid. While wax deposition may, in ure, or that if the pigging frequency for a pipeline is not correctly
fact, become negligible in the case of zero heat flux, even below the optimized, that the result will be a stuck pig and sizable production
WAT, much work would still be required to develop the technology losses (Fung et al., 2006).
required for effectively cooling the bulk fluid to this condition and Wang et al. (2008) studied the use of regular and bypass pigs in
for transporting the resulting cold slurry over long distances. Ilahi the removal of wax from pipelines in a laboratory system. The test
(2005) also discussed SINTEF and NTNU Cold Flow technology. facility used consisted of a 20 ft test section of carbon steel pipe, a
Additionally Haghighi et al. (2007) and Azarinezhad et al. (2010) mineral oil tank, a pump to push the pig with liquid as in real pig-
proposed a wet cold flow-based concept, termed HYDRAFLOW, ging operations, and a receiving tank to observe the structure of
for preventing gas hydrate agglomeration, with the potential the pigged materials. Four pressure transducers were installed to
benefit of wax inhibition. Gas hydrates are the solid solutions of monitor pressure change along the test section during pigging
gas components and water. Hammerschmidt (1934) discovered operation. Candle wax with different oil contents was cast as a film
the formation of hydrates in natural gas systems. Hydrates like or plug for measuring wax breaking force or plug transportation
waxes have concerned deep-water production at seafloor depths force, respectively. After casting, the waxy spool pieces were
of 1–3 km and temperatures between 2 and 4 °C (Gudmundsson, mounted on the test section and the pig was pushed through the
2002), conditions which encourage hydrate plug formation. Several pipe by oil from the pump, removing the wax film or plug while
studies have been done on kinetics of hydrate formation. Those the pressures at four locations along the test section were
previous studies can be categorized into two main subjects: recorded.
nucleation and growth. In contrast to previous studies, gas pipe- They concluded that the wax breaking force increases with the
lines hydrate agglomeration plays an important role. After the decrease in oil content and the increase in wax layer thickness;
break-up of the hydrate film along the interface, hydrate particles transportation force per unit plug length is affected by oil content;
agglomerate to form a hydrate plug (Lingelem et al., 1993) like transportation force decreases with the presence of oil due to
wax. Herri et al., 1999 analyzed the particle size distribution of lubrication effects; and bypass pigs exhibit a very similar breaking
hydrate particles with the particle balance equations and a mass force behaviour when compared with regular pigs. Other studies
transfer model. However it is difficult to describe agglomeration have focused on determining the optimal frequency of pigging to
from experimental observation. The particles start to agglomerate maintain a pipeline and avoid plug formation.
just after the nucleation process (Mersmann, 2002). The observed
particle size distribution is a result of kinetic contributions such 10.2. Inductive heating
as nucleation, growth, agglomeration, breakage, and attrition.
Viscosity is also a contributory factor to particle agglomeration Another possible wax removal process, studied by Sarmento et
(Mersmann, 2002). al. (2004), is the use of inductive heating of a plugged section of
pipe. They proposed this as an alternative to the use of chemicals
that react exothermically at the wax blockage to melt it, for cases
10. Wax removal methods
when the pipeline is completely blocked in a horizontal section so
that it is impossible to flow chemicals to the blockage. They tested
If wax deposition cannot be prevented, then it is imperative to
this method using the experimental setup shown in Fig. 22. They
regularly remove accumulated wax from the inside of pipeline
found that the steel layers which compose commercial flexible
walls in order to prevent the total blockage of the line. Several
lines can be heated by induction and the heat transferred to a solid
methods have thus been developed for the removal of wax depos-
wax plug in the interior of the line. They also found that their
its, including complete blockages of pipelines. Traditional methods
mathematical model, which agreed well with available experimen-
of wax removal in the petroleum industry have always had prob-
tal results, suggested that the power levels required for large-scale
lems and limitations, and they include mechanical removal, the
inductive heating might be feasible for removing wax blockage in
use of bottom hole heaters, the use of exothermic reactions such
field applications with undersea pipelines.
as that between magnesium bars and hydrochloric acid, and the
use of paraffin solvents (Woo et al., 1984). Research continues to
10.3. Biological treatment
be done to find the most efficient, cost-effective and safe methods
of removing wax deposits and blockages. Furthermore, some
Biological wax removal methods have also been studied in re-
researchers have worked on modelling the operating conditions
cent years by researchers such as Rana et al. (2010), who developed
necessary for the successful and safe restart of gelled pipelines,
systems of paraffin-degrading bacterial consortiums with nutrient
in which gelled waxy crude needs to be displaced using applied
supplements and growth enhancers for controlling paraffin deposi-
pressure.
tion in the tubular and well bore region and in surface flow lines.
Their results showed that their systems were highly effective,
10.1. Pigging eliminating the need for repeated scrapings of wax over a period
of several months. These methods are especially important be-
The practice of pigging is a way in which wax removal is com- cause, if successfully implemented, they have the benefit of provid-
monly accomplished in the field. With this method, deposited wax ing continuous control of wax deposition in pipelines through
is mechanically removed by launching a pipeline pig along the line constant biodegradation, rather than just providing a very tempo-
to scrape wax from the walls as it is forced along by the oil pres- rary fix.
sure. This, however, poses the risk of forming a wax plug down- Etoumi et al. (2008) studied the use of Pseudomonas bacteria for
stream from the pig as the scraped wax accumulates and is the reduction of wax precipitation in waxy crude oils. Their results
688 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Fig. 22. Schematic view of experimental test section for wax removal by inductive heating (Sarmento et al., 2004).

showed the ability of Pseudomonas species to emulsify immiscible as the critical shear stress value for determining whether the start
hydrocarbons such as kerosene, toluene, xylene and crude oil, an of a flow from a rest state will occur. Furthermore, they defined the
effect also studied by others, such as Sifour et al. (2007). The ob- dynamic yield stress, sd, as the parameter for describing the rela-
served overall effect of Pseudomonas treatment on crude oil tionship between shear stress and shear rate in a flow state after
showed a reduction in the concentration of long-chain hydrocar- yielding. However, the description of this yielding behaviour has
bons (C22+). Etoumi et al. concluded that Pseudomonas species seen many variations and disagreements among different authors.
may be an efficient species for reducing paraffin deposition, and Many studies have been published regarding the rheology of
that the speed of the biochemical action on crude oil is faster with- waxy crudes and their gels, the dependence of gel properties on
in the first 7 days. They also concluded that an observed reduction shear and thermal histories, and how they yield (Wardhaugh
in viscosity and WAT is indicative of the conversion of long-chain et al., 1988; Chang et al., 1998, 2000; Lopes-da-Silva and Coutinho,
alkenes to short ones. 2007; Lee et al., 2008; Oh et al., 2009). Wardhaugh and Boger
Additionally, He et al. (2003) determined through field tests, (1991), for instance, defined yield stress as ‘‘the shear stress at
that two Bacillus species and a Pseudomonas species showed good which the gelled oil ceases to behave as a Hookean solid,’’ and re-
paraffin removal properties in test wells, increasing oil production ferred to bulk yielding phenomena, when gross yielding behaviour
and eliminating the need for more expensive wax removal pro- is observed, as the yielding stress or yielding point. Houwink
cesses. Thus, biological wax removal methods may prove to be (1958) described a transition from elastic behaviour to plastic
quite effective and economically beneficial and warrant further behaviour and then to viscous flow, distinguished by a lower and
study. If a biological system can be successfully and cheaply ap- a higher yield stress. Meanwhile some researchers, such as Barnes
plied under the conditions in subsea pipelines then it will provide (1999), who noted the high degree of variation in the definition of
an extremely effective method of controlling wax deposition. yield stress, maintained that no real yield stress exists, even for
very non-Newtonian liquids. They argue this because these liquids
continue to flow or creep even below an apparent yield stress.
11. Restart of gelled pipelines
Barnes notes, however, that the concept of a yield stress is useful
for describing behaviour over a limited range.
In subsea pipelines carrying waxy crude oils that have to be
shut down temporarily for operational or emergency reasons, the
oil will eventually cool below its gel and pour points resulting in 11.1. Time-dependent gel degradation
the formation of a gel throughout the pipeline consisting of precip-
itated wax in a viscous matrix (Chang et al., 1999). This occurrence Time-dependent gel-degradation is one of the important com-
complicates the restart procedure, as the gelled oil would need to plications in modelling the restart of gelled pipelines. Ongoing ef-
be displaced in order to resume normal operations. Numerous forts to model the time-dependent rheology of gels, in order to be
researchers have addressed this problem, including Smith and able to model gel breakdown under stress, draw on research such
Ramsden (1978), Chang et al. (1999), Davidson et al. (2004), as that of Cheng and Evans (1965), Petrellis and Flumerfelt (1973),
Frigaard et al. (2007), and Vinay et al. (2007). Chang et al. (1999) and Rao et al. (1985). Recent studies of the time-dependent rheo-
modelled the isothermal restart of gelled pipelines by the applica- logical behaviour and breakdown of wax–oil gels include that done
tion of higher than normal operating pressures. In this start-up by Paso et al. (2009b), in which the mechanical behaviour of a
scenario, oil is pumped into the gelled line at high enough sus- model wax–oil gel was examined under various shear rates. Paso
tained pressure to overcome the static yield stress of the gel, thus et al. observed a convergence of shear stress values for different
breaking up the blockage and clearing the line. shear rates at absolute strain magnitudes greater than 0.1, as
The viscoplastic nature of waxy crude oils and their time- shown in Fig. 23. This was indicative of gel strength following a
dependent behaviour complicate modelling. In order to describe path-independent function of the absolute strain imposed on the
the breakdown of the gel structure along with a decrease in viscos- gel, and further indicated that the gel structure is a point function
ity, Chang et al. (1999) defined the static yield stress of the gel, ss, of the absolute strain. Based on this they concluded that, in
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 689

their model wax–oil gel at constant shear rate conditions. They


concluded that a well-defined mechanism controls the rupture of
wax crystal–crystal network linkages, and that the rheological
modelling framework based on the structural parameter, k,
provides an appropriate physical representation of the breakage
process, even for crude oils in the field with a variety of hydrocar-
bon and additive components that may cause a deviation from
third order degradation kinetics. They also proposed a model for
describing shear stress responses associated with changing shear
rates during gel degradation by applying a time-dependent
Bingham constitutive equation to experimental stress–strain data
obtained while increasing the shear rate.

11.2. Examples of restart models


Fig. 23. Measured shear stress during breakage of a wax–oil gel at shear rates
ranging from 105 s1 to 1 s1 (Paso et al., 2009b). Chang et al. (1999) went onto use a three yield stress model
proposed by Kraynik (1990), which added a dynamic yield stress
for describing behaviour after yielding to the model put forward
modelling the breakdown of a wax gel at low shear rates, the entire by Houwink (1958). The three yield stress model utilized an elas-
shear history could be represented by a single dimensionless vari- tic-limit yield stress, se, described as denoting the materials limit
able in the form of the absolute strain. of reversibility; a static yield stress, ss, described as the minimum
Paso et al. (2009b), furthermore, determined that the maximum shear stress required to cause the deformation of a material that
shear stress did not provide a useful parameter to characterize the may be described as yielding; and a dynamic yield stress, sd, de-
gel structure. Thus, in order to define a structural parameter, k, rep- scribed as the shear stress at zero shear rate, extrapolated from
resenting the fraction of unbroken crystal–crystal linkages remain- the flow curve. Chang et al. used this model to describe the three
ing in the gel structure at a given shear stress, they did so in terms possible outcomes of applying constant pressure to a gelled pipe-
of the experimental stress near the convergence point. They then line in terms of the relationship between the wall shear stress,
used this structural parameter in an nth order degradation model sw, applied to the pipeline and the initial gel strength of the oil:
to describe the gel breakage model, as shown in following
equation.  Start-up without delay (sw > ss) – Flow begins immediately with
three different regions, as shown in Fig. 25, where R is the total
1 1 radius of the pipeline and rf and rc denote the boundaries of the
ðk  ke Þ1n  ðk0  ke Þ1n ¼ ac_ b t ð18Þ
1n 1n regions:
– Flow area – The outermost region (R > r > rf), consisting of a
Here k0 and ke are the initial and equilibrium structural parameter
sheared annulus. Local stress is higher than the static yield
values, and the degradation rate parameters, n and ac_ b , were deter-
stress (s(r) > ss). The gel structure in this region is immedi-
mined by fitting experimental values of k to equation (18) via a least
ately broken down and the oil becomes liquid-like, display-
squares minimization procedure, with ke assumed to be 2  103.
ing a dynamic yield stress.
Paso et al. (2009b) were able to obtain good model fits to exper-
– Creep area – Middle region (rf > r > rc). Local stress is lower
imental values, as shown in Fig. 24. Their fitted degradation order
than static yield stress, but higher than elastic-limit yield
for different shear rates ranged from 2.7 to 3.33, indicating that a
stress (ss > s > se). Gel structure in this region begins to
third order degradation mechanism controls the breakdown of
degrade with a viscoelastic deformation.
– Elastic deformation area – Innermost region (r < rc). Local
stress is lower than elastic-limit yield stress (s < se). Solid-
like core where oil only undergoes elastic deformation. Will
initially move with creep region as an unsheared plug of
radius, r, until the gel in the creep region degrades from
the outside in, leaving only the core as the plug.
 Start-up with delay (ss > sw > se) – Flow begins after a delay
time, tdelay. Exterior creep region and interior elastic deforma-
tion area exist and, initially, no flow occurs. Flow only begins
once gel in the creep region has sufficiently degraded, starting
at the wall, allowing for movement of an unsheared plug with
uniform velocity through the pipe. The size of the plug (r) will
decrease as degradation in the creep region continues.
 Unsuccessful start-up (sw < se) – Flow will not start under this
condition. Oil only deforms elastically and gel structure is unaf-
fected by shear.

Chang et al. (1999) noted that for a successful start-up, the


gelled oil in a cross-section of pipe will become heterogeneous be-
cause of differences in the rate of structural breakdown caused by
differences in local shear stress. Therefore, their model takes into
Fig. 24. Comparison of experimental and fitted k values at a shear rate of 103 s1.
account the time-dependent rheology of the waxy crude oils. A
The optimized reaction rate order is 3.07, with a rate constant of 0.131 s1 (Paso time-dependent Bingham-style equation, shown in Eqs. (19a)-
et al., 2009b). (19c) was used for an approximation of the time-dependent,
690 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Fig. 25. Schematic diagram of start-up without delay (sw>ss when t = 0)(Chang et al., 1999).

non-Newtonian behaviour of a gelled waxy crude oil under con- For the case of start-up without delay (or start-up with delay at
trolled stress conditions. time, t > tdelay), Chang et al. (1999) define the initial wall shear
stress in terms of the pressure drop. To model the time- and posi-
s ¼ sy ðtÞ þ gðtÞc_ ; s > sy ðtÞ ð19aÞ tion-dependent changes in the flow properties of the OGF, Chang
et al. (1999) used a finite differences method. M time intervals
sy ð0Þ  sy ð1Þ were used to divide the duration of the flow from start-up
sy ðtÞ ¼ þ sy ð1Þ ð19bÞ
1 þ kt (Dt = ti–ti1), and the flow was treated as approximately steady
in each time interval for sufficiently small Dt. The sheared annulus
gðtÞ ¼ constant ð19cÞ (r < r R) was divided, for each instant, ti, into N radial elements of
Here g is the plastic viscosity, c_ is the shear rate, sy is the appar- thickness, Dr, and distance, rj, from the centre of the pipe
ent yield stress governing the behaviour of the oil, and k is a rate (rj = rj1 + Dr = r + jDr). The volumetric flow rate, Qi, at time ti
constant. sy(0) and sy(1) are the apparent yield stress at times was thus given by following equation,
t = 0 and t = 1, respectively. sy(0) would be equivalent to ss(0), X
N
with an initial wall stress above this value resulting in an instanta- Qi ¼ Q j þ Q plug ð21Þ
neous finite flow rate, and sy(1) coincides with se(0), with wall j¼1

stresses below this value resulting in reversible deformation and where Qplug is the flow rate of the unsheared plug and Qj is the
no possible flow. volumetric flow rate of the jth annular element. In a later work,
The basic physical model used by Chang et al. (1999) to describe However, Davidson et al. (2004) maintained that the finite differ-
the start-up process was the pumping of an incoming fluid (ICF) ences method was unnecessary, because of the quasi-steady state
into a pipe of length, L, and inside diameter, D, to displace the out- assumption for the OGF, which was represented as a Bingham fluid
going fluid (OGF), as shown in Fig. 26. Here Z(t) is the length of the that would have apparent yield stress and plastic viscosity inde-
pipe occupied by the ICF, r I ðtÞ and r o ðtÞ are the unsheared plug ra- pendent of the shear rate and thus the radial position.
dii of the ICF and OGF respectively at time, t, P1 is the inlet pres- Using their model, Chang et al. (1999) could calculate the plug ra-
sure, P2 is the exit pressure, and Pz is the interface pressure. The dius, r, for each time interval using equations (19b) and (20) with a
radius of the unsheared plug in the flow was given by equation, known wall stress. The flow rate could then be computed from j = N
sy ðtÞ at the pipe wall inward to the unsheared plug at j = 0. The onset of
r  ðtÞ ¼ R ð20Þ turbulence was predicted by calculation of a critical Reynolds num-
swo ðtÞ
ber, with the appropriate adjustment to the friction factor used in
where swo(t) is the wall shear stress in the OGF at time t. the model. For these calculations the pipe dimensions (L and R),

Fig. 26. Schematic diagram of two-fluid displacement model. (a) True interface; (b) simplified interface (Chang et al., 1999).
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 691

pump pressure (DPc), properties of the OGF (sy(0), sy(1), g(t), k and where k is the current subdivision out of M initial subdivisions of
qo), and properties of the ICF (sB, gB and qI) need to be known. Here qI the gelled oil used in the calculations; and qk is the dimensionless
and qo are the fluid densities of the ICF and OGF; sB is the Bingham density and Q k the dimensionless volumetric flow rate of the cur-
yield stress; and gB is the Bingham plastic viscosity. Therefore, the rent subdivision. The pressure drop over each segment was calcu-
accuracy of this model in predicting the time-dependent flow prop- lated using following equation.
erties during start-up and the time needed to clear a blockage de-
Pk fk qk t2k
pended greatly on comprehensive knowledge of the system, swk ¼ ¼ ð24Þ
requiring accurate experimental measurements. 4DLk 2
Davidson et al. (2004) developed another model for the restart of
In the calculation procedure used by Davidson et al. (2004), the
gelled pipelines. This model extended the one developed by Chang
length and density of each segment of oil is then updated to
et al. (1999) to account for the compressibility and inhomogeneity
account for the displacement of the OGF from the length of pipe
of the gelled oil and displacing fluid. In this model, when the inlet
and the increase in the length of the ICF within the pipe. The ICF
pressure creating a wall stress in excess of the static yield stress is
rheology is assumed to be time-independent and it is therefore
applied to the gelled oil, initially only a narrow region of length Lf de-
separated into segments of equal length in each time step, with
forms and breaks down under stress. This yielded region is com-
the number of segments increasing by one with each time step.
pressed by the entering ICF, which is also compressed. Eventually,
The length of an ICF segment, DLICF, in time interval, i, at time
at time t = t0 the entire gelled oil plug will yield and move together
t  t 0 is given by following equation.
with the ICF at the same mass flow rate, as shown in Fig. 27.
For the calculations in this model from Davidson et al. (2004), P
LICF DL  m OGF
k¼1 Lk
the bulk mass flow rate, G, is first guessed (can use value from pre- DLICF ¼ ¼ ð25Þ
K þi K þi
vious time step). Then the frictional factor, fk, is calculated by iter-
ation for each longitudinal ICF and OGF segment at current time Here K is the number of ICF segments chosen at time t = t0, m is
step using the Buckingham–Reiner equation for pipe flow of a the number of remaining OGF segments within the pipe, and DLOGF k
time-independent Bingham fluid, and empirical relationships is the length of the k th gelled oil segment in the OGF for time
developed by other authors for calculating the frictional factor in t  t 0 . Davidson et al. (2004) also calculated DLOGF
k and the average
laminar and turbulent flow. Equations (22) and (23) are used to density for the k th segment in dimensionless form. Next in their
evaluate the mean velocity and shearing time, tsk, calculation procedure, the location of each segment and the ICF–
qk Q k ¼ G ð22Þ OGF interface is determined relative to the downstream end of
the OGF plug. Then the pressure drop over each ICF and OGF
k1 segment is summed to give the overall pressure drop, and the
t sk ðtÞ ¼ t  t0 ð23Þ difference between this value and the applied pressure drop
M1

Fig. 27. Schematic of compression flow (Davidson et al., 2004).


692 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

calculated and mass flow rate adjusted accordingly. This overall Azarinezhad, R., Chapoy, A., Anderson, R., Tohidi, B., 2010. A wet cold-flow
technology for tackling offshore flow-assurance. SPE Projects, Facilities and
process is iterated until the difference is negligible. G iterates to
Construction 5, 58–64.
zero in the case that the applied pressure is not high enough to Azevedo, L.F.A., Teixeira, A.M., 2003. A critical review of the modeling of wax
start flow at a given time, and the calculated pressure drop be- deposition mechanisms. Petrol. Sci. Technol. 21, 393–408.
comes the minimum required for start-up. The results of this mod- Bagatin, R., Busto, C., Correra, S., Margarone, M., Carniani, C., 2008. Wax modelling:
there is need for alternatives. In: SPE Russian Oil & Gas Technical Conference
el were significantly different from those of the earlier model and Exhibition. Moscow, 2008. Society of Petroleum Engineers.
developed by Chang et al. (1999), indicating the importance of fully Banki, R., Hoteit, H., Firoozabadi, A., 2008. Mathematical formulation and numerical
understanding the mechanism by which the gelled oil yields and is modeling of wax deposition in pipelines from enthalpy–porosity approach and
irreversible thermodynamics. Int. J. Heat Mass Transfer 51, 3387–3398.
displaced, and of determining the most realistic assumptions that Barnes, H.A., 1999. The Yield stress—a review or ‘pamsa qei’—everything flows? J.
can be made during modelling. Non-Newtonian Fluid Mech. 81, 133–178.
Other researchers have also tackled understanding, modelling Bello, O.O., Fasesan, S.O., Teodoriu, C., Reinicke, K.M., 2006. An evaluation of the
performance of selected wax inhibitors on paraffin deposition of Nigerian crude
and optimizing the restart of gelled lines. Borghi et al. (2003) oils. Pet. Sci. Technol. 24, 195–206.
developed a model focusing on solid-like fracture propagation, vis- Bilderback, C.A., McDougall, L.A., 1963. Complete paraffin control in petroleum
cous dissipation and compression of the broken gelled oil. Ekwer- production. SPE J. Petrol. Technol. 21, 1151–1156.
Borghi, G.P., Correra, S., Merlini, M., Carniani, C., 2003. Prediction and scaleup of
ibe et al. (2009), for instance, studied the effect of system pressure waxy oil restart behavior. In: SPE International Symposium on Oilfield
on the restart of gelled subsea pipelines. They determined that Chemistry. Houston, 2003. Society of Petroleum Engineers.
higher system pressures in subsea pipelines could lead to the for- Bummer, B.L., 1971. Improved paraffin prevention techniques reduce operating
costs, powder river basin, wyoming. In: Rocky Mountain Regional Meeting of
mation of a weaker gel with lower yield strength, which would
the Society of Petroleum Engineers of AIME. Billings, 1971. American Institute
mean that the necessary applied pressure for displacing it would of Mining, Metallurgical, and Petroleum Engineers.
be more easily and cheaply achieved than might be predicted. Burger, E.D., Perkins, T.K., Striegler, J.H., 1981. Studies of wax deposition in the trans
alaska pipeline. J. Petrol. Technol. 33, 1075–1086.
Carmen García, M., Urbina, A., 2003. Effect of crude oil composition and blending on
flowing properties. Petrol. Sci. Technol. 21, 863–878.
12. Conclusions Carmen García, M., 2001. Paraffin deposition in oil production. In: SPE International
Symposium on Oilfield Chemistry. Houston, 2001. Society of Petroleum
Engineers.
Contention still remains as to the specific mechanisms that gov- Carmen García, M., Orea, M., Carbognani, L., Urbina, A., 2001. The effect of paraffinic
ern wax deposition in pipelines. However, the importance of fractions on crude oil wax crystallization. Petrol. Sci. Technol. 19, 189–196.
molecular diffusion is generally accepted and shear dispersion is Chakrabarti, D.P., Das, G., Das, P.K., 2006. The transition from water continuous to
oil continuous flow pattern. AIChE J. 52, 3668–3678.
usually not dismissed, at least due to the involvement of shear Chakrabarti, D.P., Das, G., Das, P.K., 2007. Identification of stratified liquid–liquid
forces in the removal of wax deposits, the accounting of which flow through horizontal pipes by a non-intrusive optical probe. Chem. Eng. Sci.
has been shown by some authors to have a great impact on the 62, 1861–1876.
Chang, C., Boger, D.V., Nguyen, Q.D., 1998. The yielding of waxy crude oils. Ind. Eng.
accuracy of wax deposition models. Many models have been devel- Chem. Res. 37, 1551–1559.
oped based on the importance of these mechanisms, for which the Chang, C., Boger, D.V., Nguyen, Q.D., 2000. Influence of thermal history on the waxy
approach to a realistic representation of the solid phase wax com- structure of statically cooled waxy crude oil. SPE J. 5, 148–157.
Chang, C., Nguyen, Q.D., Rønningsen, H.P., 1999. Isothermal start-up of pipeline
ponents has a significant impact on accuracy. Recently, a correct transporting waxy crude oil. J. Non-Newtonian Fluid Mech. 87, 127–154.
heat-mass transfer analogy has been introduced into the modelling Chen, X., Tsang, Y., Zhang, H.Q., Chen, T.X., 2007. Pressure-wave propagation
of wax deposition, allowing for more accurate prediction across the technique for blockage detection in subsea flowlines. In: SPE Annual
Technical Conference and Exhibition. Anaheim, 2007. Society of Petroleum
range of possible precipitation kinetics. In the future even more
Engineers.
accurate and robust models will be possible by combining this Chen, X.T., Butler, T., Volk, M., Brill, J.P., 1997. Techniques for measuring wax
new approach with an increased understanding of the mechanisms thickness during single and multiphase flow. In: SPE Annual Technical
Conference and Exhibition. San Antonio, 1997. Society of Petroleum
involved in wax deposition and gelation and of the impact of other
Engineers.
species present in crude oil, such as asphaltenes and emulsified Cheng, D.C.H., Evans, F., 1965. Phenomenological characterization of the rheological
water. behavior of inelastic reversible thixotropic and antithixotropic fluids. Brit. J.
Understanding wax aging mechanisms is also very important to Appl. Phys. 16, 1599–1617.
Cordoba, A.J., Schall, C.A., 2001. Application of a heat method to determine wax
fully understanding the process of the formation of wax deposits in deposition in a hydrocarbon binary mixture. Fuel 80, 1285–1291.
pipelines. Furthermore, understanding these mechanisms and pre- Correra, S., Fasano, A., Fusi, L., Merino-Garcia, D., 2007. Calculating deposit
dicting the CCN of particular crude oils would be helpful in deter- formation in the pipelining of waxy crude oils. Meccanica 42, 149–165.
Coutinho, J.A.P., Daridon, J.L., 2005. The limitations of the cloud point measurement
mining what chemical inhibitors would be most effective for techniques and the influence of the oil composition on its detection. Pet. Sci.
preventing wax build-up in pipelines carrying those oils. The Technol. 23, 1113–1128.
continuing research into methods of inhibiting wax deposition Coutinho, J.A.P., 1998. Predictive UNIQAC: a new model for the description of
multiphase solid-liquid equilibria in complex hydrocarbon mixtures. Ind. Eng.
and removing deposits has the potential of making the mainte- Chem. Res. 37, 4870.
nance of crude oil pipelines significantly easier, as it becomes eas- Coutinho, J.A.P., Lopes-da-Silva, J.A.L., Ferreira, A., Soares, M.S., Daridon, J.L., 2003.
ier to optimize pigging frequency, to determine the minimum Evidence for the aging of wax deposits in crude oils by ostwald ripening. Petrol.
Sci. Technol. 21, 381–391.
pressure required to restart gelled lines, or even to avoid the need
Davidson, M.R., Nguyen, Q.D., Chang, C., Rønningsen, H.P., 2004. A model for restart
for constant wax removal procedures by finding a way to cost- of a pipeline with compressible gelled waxy crude oil. J. Non-Newtonian Fluid
effectively implement a promising method of control such as the Mech. 123, 269–280.
de Oliveira, M.C.K., Carvalho, R.M., Carvalho, A.B., Couto, B.C., Faria, F.R.D., Cardoso,
use of polar crude oil fractions or biological removal measures.
R.L.P., 2010. Waxy crude oil emulsion gel: impact on flow assurance. Energy
Fuels 24, 2287–2293.
Duffy, D.M., Rodger, P.M., 2002. Modeling the activity of wax inhibitors: a case
References study of poly(octadecyl acrylate). J. Phys. Chem. B 106, 11210–11217.
Duffy, D.M., Moon, C., Rodger, P.M., 2004. Computer-assisted design of oil additives:
hydrate and wax inhibitors. Mol. Phys. 102, 203–210.
Addison, G.E., 1984. Paraffin Control More Cost-Effective. In: SPE Eastern Regional
Edmonds, B., Moorwood, T., Szczepanski, R., Zhang, X., 2008. Simulating wax
Meeting. Charleston, 1984. Society of Petroleum Engineers.
deposition in pipelines for flow assurance. Energy Fuels 22, 729–741.
Al-Sabagh, A.M., El-Kafrawy, A.F., Khidr, T.T., El-Ghazawy, R.A., Mishrif, M.R., 2007.
Ekweribe, C.K., 2008. Quiescent Gelation of Waxy Crudes and Restart of Shut-in
Synthesis and evaluation of some novel polymeric surfactants based on
Subsea Pipelines (MS Thesis). Norman, Oklahoma: University of Oklahoma.
aromatic amines used as wax dispersant for waxy gas oil. J. Dispersion Sci.
Ekweribe, C.K., Civan, F., Lee, H.S., Singh, P., 2009. Interim Report on Pressure Effect
Technol. 28, 976–983.
on Waxy-Crude Pipeline-Restart Conditions Investigated by a Model System.
Asperger, R.G., Sattler, R.E., Tolonen, W.J., Pitchford, A.C., 1981. Prediction of Wax
SPE Projects, Facilities & Construction, pp. 61–74.
Buildup in 24 inch, Cold Deep Sea Oil Loading Line. USMS 10363.
A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694 693

Etoumi, A., El Musrati, I., El Gammoudi, B., El Behlil, M., 2008. The reduction of wax Lira-Galeana, C., Firoozabadi, A., Prausnitz, J.M., 1996. Thermodynamics of wax
precipitation in waxy crude oils by Pseudomonas species. J. Ind. Microbiol. precipitation in petroleum mixtures. AIChE J. 42, 239–248.
Biotechnol. 35, 1241–1245. Lockhart, T., Correra, S., 2005. Colloid and Surface Science Issues in the Petroleum
Farina, A., Fasano, A., 1997. Flow characteristics of waxy crude oils in laboratory Industry. In: Ente Nazionale Idrocarburiz & Istituto della Enciclopedia Italiana
experimental loops. Math. Comput. Model. 25, 75–86. Encyclopaedia of Hydrocarbons. Roma, Italy: Istituto della Enciclopedia Italiana.
Fasano, A., Fusi, L., Correra, S., 2004. Mathematical models for waxy crude oils. Ch. 3.2.1. pp. 167–177.
Meccanica 39, 441–482. Lopes-da-Silva, J.A., Coutinho, J.A.P., 2007. Analysis of the isothermal structure
Fielder, M., Johnson, R.W., 1986. The use of pour-point depressant additive in the development in waxy crude oils under quiescent conditions. Energy Fuels 21,
beatrice field. In: SPE European Petroleum Conference. London, 1986. Society of 3612–3617.
Petroleum Engineers. Majeed, A., Bringedal, B., Overa, S., 1990. Model calculates wax deposition for n. sea
Frigaard, I., Vinay, G., Wachs, A., 2007. Compressible displacement of waxy oils. Oil Gas J. 88, 63–69.
crude oils in long pipeline startup flows. J. Non-Newtonian Fluid Mech. Manka, J.S., Ziegler, K.L., 2001. Factors affecting the performance of crude oil wax-
147, 45–64. control additives. In: SPE Production and Operations Symposium. Oklahoma,
Fulford, R.S., 1975. Oilwell paraffin prevention chemicals. In: SPE Regional Meeting. 2001. Society of Petroleum Engineers.
Oklahoma, 1975. American Institute of Mining, Metallurgical, and Petroleum Manka, J.S., Magyar, J.S., Smith, R.P., 1999. A novel method to winterize traditional
Engineers. pour point depressants. In: SPE Annual Technical Conference and Exhibition.
Fung, G., Backhaus, W.P., McDaniel, S., Erdogmus, M., 2006. To pig or not to pig: the Houston, 1999. Society of Petroleum Engineers.
marlin experience with stuck pig. In: Offshore Technology Conference. Houston, Matzain, A., 1999. Multiphase Flow Paraffin Deposition Modeling (PhD
2006. Offshore Technology Conference. Dissertation). Tulsa, Oklahoma: The University of Tulsa.
Fusi, L., Farina, A., 2004. A mathematical model for bingham-like fluids with visco- Matzain, A., Apte, M.S., Zhang, H., Volk, M., Brill, J.P., Creek, J.L., 2002. Investigation
elastic core. Z. Angew. Math. Phys. 55, 826–847. of paraffin deposition during multiphase flow in pipelines and wellbores—part
Fusi, L., 2003. On the stationary flow of a waxy crude oil with deposition 1: experiments. J. Energy Res. Technol. 124, 180–186.
mechanisms. Nonlinear Anal. 53, 507–526. Mendell, J.L., Jessen, F.W., 1970. Mechanism of Inhibition of Paraffin Deposition in
Groffe, D., Groffe, P., Takhar, S., Andersen, S.I., Stenby, E.H., Lindeloff, N. et al., 2001. Crude Oil Systems. In: Ninth Bienniel Production Techniques Symposium.
A wax inhibition solution to problematic fields: a chemical remediation process. Wichita Falls, 1970. American Institute of Mining, Metallurgical and Petroleum
Petrol. Sci. Technol. 19, 205–217. Engineers.
Gudmundsson, J.S., ‘‘Cold Flow Technology’’. In: Proc. 4th Intnl. Hydrates Conf., Merino-Garcia, D., Correra, S., 2008. Cold flow: a review of a technology to avoid
Yokohama (2002), pp. 912–916. wax deposition. Pet. Sci. Technol. 26, 446–459.
Haghighi, H., Azarinezhad, R., Chapoy, A., Anderson, R., Tohidi, B., 2007. Hydraflow: Merino-Garcia, D., Margarone, M., Correra, S., 2007. Kinetics of waxy gel formation
avoiding gas hydrate problems. In: SPE Europe/EAGE Annual Conference and from batch experiments. Energy Fuels 21, 1287–1295.
Exhibition. London, 2007. Society of Petroleum Engineers. Mersmann, A. 2002. Crystallization Technology Handbook, second ed. Marcel
Hammami, A., Ratulowski, J., Coutinho, J.A.P., 2003. Cloud points: can we measure Dekker.
or model them? Petrol. Sci. Technol. 21, 345–358. Nautiyal, S.P., Kumar, S., Srivastava, S.P., 2008. Crystal structure of n-paraffin
Hammerschmidt, E.G., 1934. Formation of gas hydrates in natural gas concentrates of crude oils. Pet. Sci. Technol. 26, 1339–1346.
transportation lines. Ind. Eng. Chem. 26, 851–855. Newberry, M.E., Barker, K.M., 1985. Formation damage prevention through the
Haq, M.A., 1978. Deposition of Paraffin Wax from its Solution with Hydrocarbons control of paraffin and asphaltene deposition. In: SPE Production Operations
(USMS 10541). Society of Petroleum Engineers. Symposium. Oklahoma City, 1985. Society of Petroleum Engineers.
He, Z., Mei, B., Wang, W., Sheng, J., Zhu, S., Wang, L. et al., 2003. A pilot test using Oh, K., Jemmett, M., Deo, M., 2009. Yield behavior of gelled waxy oil: effect of stress
microbial paraffin-removal technology in liaohe oilfield. Petrol. Sci. Technol. 21, application in creep ranges. Ind. Eng. Chem. Res. 48, 8950–8953.
201–210. Paso, K.G., Fogler, H.S., 2003. Influence of n-paraffin composition on the aging of
Hernandez, O.C., Hensley, H., Sarica, C., Brill, J.P., Volk, M., Delle-Case, E., 2004. wax-oil gel deposits. AIChE J. 49, 3241–3252.
Improvements in single-phase paraffin deposition modeling. SPE Prod. Oper. 19, Paso, K.G., 2005. Paraffin Gelation Kinetics (PhD Dissertation). Ann Arbor, Michigan:
237–244. University of Michigan.
Herri, J.-M., Pic, J.S., Gruy, F., Cournil, M., 1999. Methane hydrate crystallization Paso, K.G., Kompalla, T., Aske, N., Rønningsen, H.P., Øye, G., Sjöblom, J., 2009a. Novel
mechanism from in-situ particle size analyzer. AIChE J. 45, 590–602. surfaces with applicability for preventing wax deposition: a review. J.
Holder, G.A., Winkler, J., 1965. Wax crystallization from distillate fuels. J. Inst. Petrol. Dispersion Sci. Technol. 30, 757–781.
51, 228–252. Paso, K.G., Kompalla, T., Oschmann, H.J., Sjöblom, J., 2009b. Rheological degradation
Houwink, R., 1958. Elasticity, Plasticity and Structure of Matter. Cambridge of model wax–oil gels. J. Dispersion Sci. Technol. 30, 472–480.
University Press, New York. Paso, K.G., Silset, A., Sørland, G., Gonçalves, M.A.L., Sjöblom, J., 2009c.
Hunt, E.B.J., 1962. Laboratory study of paraffin deposition. SPE J. Petrol. Technol. 14, Characterization of the formation, flow ability, and resolution of Brazilian
1259–1269. crude oil emulsions. Energy Fuels 23, 471–480.
Ilahi, M., 2005. Evaluation of Cold Flow Concepts (MS Thesis). Trondheim, Norway: Patton, C.C., 1970. paraffin deposition from refined wax-solvent systems. Soc.
Norwegian University of Science and Technology. Petrol. Eng. J., 17–24.
Jang, Y.H., Blanco, M., Creek, J., Tang, Y., Goddard, W.A., 2007. Wax Inhibition Pedersen, K.S., Rønningsen, H.P., 2003. Influence of wax inhibitors on wax
by comb-like polymers: support of the incorporation-perturbation appearance temperature, pour point, and viscosity of waxy crude oils. Energy
mechanism from molecular dynamics simulations. J. Phys. Chem. B 111, Fuels 17, 321–328.
13173–13179. Petrellis, N.C., Flumerfelt, R.W., 1973. Rheological behavior of shear degradable oils:
Jennings, D.W., Breitigam, J., 2009. Paraffin inhibitor formulations for different kinetic and equilibrium properties. Can. J. Chem. Eng. 51, 291–301.
application environments: from heated injection in the desert to extreme cold Prausnitz, J.M., Lichtenthaler, R.N., Azevedo, E.G., 1986. Molecular Thermodynamics
arctic temperatures. Energy Fuels 24, 2337–2349. of Fluid-phase Equilibria. Prentice-Hall, Englewood Cliffs, NJ.
Jorda, R.M., 1966. Paraffin deposition and prevention in oil wells. SPE J. Petrol. Quenelle, A., Gunaltun, M., 1987. Comparison between thermal insulation coatings
Technol. 18, 1605–1612. for underwater pipelines. In: Offshore Technology Conference. Houston, 1987.
Kelland, M.A., 2009. Production Chemicals for the Oil and Gas Industry. CRC Offshore Technology Conference.
Press. Raj, T.S., Chakrabarti, D.P., Das, G., 2005. Liquid-liquid stratified flow through
Kraynik, A.M., 1990. ER fluid standards: comments on er fluid rheology. In: Carlson, horizontal conduits. Chem. Eng. Technol. 28, 899–907.
J.D., Sprecher, A.F., Conrad, H. (Eds.), Proceedings of the Second International Ramírez-Jaramillo, E., Lira-Galeana, C., Manero, O., 2001. Numerical Simulation of
Conference on ER Fluids. Raleigh, 1990. Technomic Publishing Company. Wax Deposition in Oil Pipeline Systems. Petrol. Sci. Technol. 19, 143–156.
Kriz, P., Andersen, S.I., 2005. Effect of asphaltenes on crude oil wax crystallization. Ramírez-Jaramillo, E., Lira-Galeana, C., Manero, O., 2004. Modeling wax deposition
Energy Fuels 19, 948–953. in pipelines. Petrol. Sci. Technol. 22, 821–861.
Lee, H.S., 2008. Computational and Rheological Study of Wax Deposition and Rana, D.P., Bateja, S., Biswas, S.K., Kumar, A., Misra, T.R., Lal, B., 2010. Novel
Gelation in Subsea Pipelines (PhD Dissertation). Ann Arbor, Michigan: microbial process for mitigating wax deposition in down hole tubular and
University of Michigan. surface flow lines. In: SPE Oil and Gas India Conference. Mumbai, 2010. Society
Lee, H.S., Singh, P., Thomason, W.H., Fogler, H.S., 2008. Waxy oil gel breaking of Petroleum Engineers.
mechanisms: adhesive versus cohesive failure. Energy Fuels 22, 480–487. Rao, B.M.A., Mahajan, S.P., Khilar, K.C., 1985. A Model on the breakdown of crude oil
Leelavanichkul, P., Deo, M.D. Hanson, F.V., 2004. Crude oil characterization and gel. Can. J. Chem. Eng. 63, 170–172.
regular solution approach to thermodynamic modeling of solid precipitation at Reistle, C.E.J., 1928. Methods of Dealing with Paraffin Troubles Encountered in
low pressure. Petrol. Sci. Technol. 22, 973–990. Producing Crude Oils. USBM Technical Papers.
Leiroz, A.T., Azevedo, L.F.A., 2005. Studies on the mechanisms of wax deposition in Reistle, C.E.J., 1932. Paraffin and Congealing Oil Problems. USBM Bulletins.
pipelines. In: Offshore Technology Conference. Houston, 2005. Offshore Sarmento, R.C., Ribbe, G.A.S., Azevedo, L.F.A., 2004. Wax blockage removal by
Technology Conference. inductive heating of subsea pipelines. Heat Transfer Eng. 25, 2–12.
Li, H., Zhang, J., Yan, D., 2005. Correlations between the pour point or gel point and Senra, M., Panacharoensawad, E., Kraiwattanawong, K., Singh, P., Fogler, H.S., 2008.
the amount of precipitated wax for waxy crudes. Pet. Sci. Technol. 23, 1313– Role of n-alkane polydispersity on the crystallization of n-alkanes from
1322. solution. Energy Fuels 22, 545–555.
Lingelem, M.N., Majeed A.I., Stange, E. Industrial experience in evaluation of hydrate Senra, M., Scholand, T., Maxey, C., Fogler, H.S., 2009. Role of polydispersity and
formation, inhibition, and dissociation in pipeline design and operation. In: cocrystallization on the gelation of long-chained n-alkanes in solution. Energy
Proc. 1st Intnl. Hydrates Conf., New-Paltz, NY (1993). Fuels 23, 5947–5957.
694 A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671–694

Sifour, M., Al-Jilawi, M.H., Aziz, G.M., 2007. Emulsification properties of Venkatesan, R., Fogler, H.S., 2004. Comments on analogies for correlated heat and
biosurfactant produced from Pseudomonas aeruginosa RB 28. Pak. J. Biol. Sci. mass transfer in turbulent flow. AIChE J. 50, 1623–1626.
10, 1331–1335. Venkatesan, R., Östlund, J., Chawla, H., Wattana, P., Nydén, M., Fogler, H.S., 2003. The
Singh, P., Venkatesan, R., Fogler, H.S., 2000. Formation and aging of incipient thin Effect of asphaltenes on the gelation of waxy oils. Energy Fuels 17, 1630–1640.
film wax–oil gels. AIChE J. 46, 1059–1074. Vinay, G., Wachs, A., Frigaard, I., 2007. Start-up transients and efficient computation
Singh, P., Venkatesan, R., Fogler, H.S., 2001a. Morphological evolution of thick wax of isothermal waxy crude oil flows. J. Non-Newtonian Fluid Mech. 143, 141–
deposits during aging. AIChE J. 47, 6–18. 156.
Singh, P., Youyen, A., Fogler, H.S., 2001b. Existence of a critical carbon number in the Visintin, R.F.G., Lockhart, T.P., Lapasin, R., D’Antona, P., 2008. Structure of waxy
aging of a wax-oil gel. AIChE J. 47, 2111–2124. crude oil emulsion gels. J. Non-Newtonian Fluid Mech. 149, 34–39.
Singhal, H.K., Sahai, G.C., Pundeer, G.S., Chandra, K., 1991. Designing and selecting Wang, K.S., Wu, C.H., Creek, J.L., Shuler, P.J., Tang, Y., 2003. Evaluation of effects of
wax crystal modifier for optimum field performance based on crude oil selected wax inhibitors on paraffin deposition. Petrol. Sci. Technol. 21, 369–379.
composition. In: Annual Technical Conference and Exhibition of the Society of Wang, Q., Sarica, C. & Volk, M., 2008. An experimental study on wax removal in
Petroleum Engineers. Dallas, 1991. Society of Petroleum Engineers. pipes with oil flow. J. Energy Resour. Technol., 130.
Smith, P.B., Ramsden, R.M.J., 1978. The prediction of oil gelation in submarine Wardhaugh, L.T., Boger, D.V., 1991. The measurement and description of the
pipelines and the pressure required for restarting flow. In: European Offshore yielding behavior of waxy crude oil. J. Rheol. 35, 1121–1156.
Petroleum Conference and Exhibition. London, 1978. European Offshore Wardhaugh, L.T., Boger, D.V., Tonner, S.P., 1988. Rheology of waxy crude oils. In: SPE
Petroleum Conference and Exhibition. International Meeting on Petroleum Engineering. Tianjin, 1988. Society of
Solaimany Nazar, A.R., Dabir, B., Islam, M.R., 2005a. A multi-solid phase Petroleum Engineers.
thermodynamic model for predicting wax precipitation in petroleum Woo, G.T., Garbis, S.J. & Gray, T.C., 1984. Long-term control of paraffin deposition.
mixtures. Energy Sources 27, 173–184. In: Technical Conference and Exhibition. Houston, 1984. Society of Petroleum
Solaimany Nazar, A.R., Dabir, B., Islam, M.R., 2005b. Experimental and mathematical Engineers.
modeling of wax deposition and propagation in pipes transporting crude oil. Wuhua, C., Zongchang, Z., 2006. Thermodynamic modeling of wax precipitation in
Energy Sources 27, 185–207. crude oils. Chin. J. Chem. Eng. 14, 685–689.
Soni, H.P., Kiranbala, Bharambe, D.P., 2008. Performance-based design of wax Yang, X., Kilpatrick, P., 2005. Asphaltenes and waxes do not interact synergistically
crystal growth inhibitors. Energy Fuels 22, 3930–3938. and coprecipitate in solid organic deposits. Energy Fuels 19, 1360–1375.
Tinsley, J.F., Prud’homme, R.K., Guo, X., Adamson, D.H., Shao, S., Amin, D. et al., 2007. Zaman, M., Agha, K.R., Islam, M.R., 2006. Laser based detection of paraffin in crude
Effects of polymers on the structure and deposition behavior of waxy oils. In: oil samples: numerical and experimental study. Pet. Sci. Technol. 24, 7–22.
SPE International Symposium on Oilfield Chemistry. Houston, 2007. Society of Zaman, M., Bjorndalen, N., Islam, M.R., 2004. Detection of precipitation in pipelines.
Petroleum Engineers. Petrol. Sci. Technol. 22, 1119–1141.
Torres, G.A., Turner, C., 2005. Method of straight lines for a bingham problem as a
model for the flow of waxy crude oils. Electron. J. Diff. Equat., 1–15.

View publication stats

You might also like