You are on page 1of 24

International Journal of Multiphase Flow 37 (2011) 671694

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
b

Department of Chemical Engineering, The University of the West Indies, Trinidad and Tobago
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

Article history:
Received 23 December 2010
Received in revised form 9 February 2011
Accepted 20 February 2011
Available online 27 February 2011
Keywords:
Waxy crude oil
Oil-pipe
Solidsolid transition
Solidliquid equilibrium
Wax precipitation
wax removal

a b s t r a c t
The gelling of waxy crudes and the deposition of wax on the inner walls of subsea crude oil pipelines
present a costly problem in the production and transportation of oil. The timely removal of deposited
wax is required to address the reduction in ow rate that it causes, as well as to avoid the eventual loss
of a pipeline in the event that it becomes completely clogged. In order to understand this problem and
address it, signicant 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
walls and of removing accumulated wax have been studied to nd the most efcient and cost-effective
means of maintaining pipelines prone to wax deposition. This paper seeks to review the current state of
research into these areas, highlighting what is so far understood about the mechanisms guiding this
wax deposition, and how this knowledge can be applied to modelling and providing solutions to this
problem.
2011 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

4.
5.
6.

7.

8.
9.

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

672
672
672
672
673
673
673
673
673
673
674
674
674
675
675
675
676
677
679
679
680
680
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 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijmultiphaseow.2011.02.007

672

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

9.1.
9.2.

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

1. Introduction
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,
1963; Haq, 1978). For subsea pipelines, in particular, it has become
especially important to solve the issue of wax build-up, as largescale oil production in colder regions will be faced with more
severe wax precipitation (Smith and Ramsden, 1978; Asperger
et al., 1981).
Wax precipitation within pipelines at and below the Cloud
Point or Wax Appearance Temperature (WAT) can lead to gelling
that inhibits ow by causing signicant non-Newtonian behaviour
and increasing effective viscosities as the temperature of a waxy
crude oil approaches its Pour Point (Pedersen and Rnningsen,
2003). Alternatively, when just the pipeline wall is below the
WAT, this promotes the deposition of a layer of parafn molecules
that can grow over time, constricting ow. This is especially problematic for pipelines in deep-sea environments, as, even in relatively warm climates, the water temperature will be on the order
of 5 C (Azevedo and Teixeira, 2003).
Some researchers, such as Carmen Garca et al. (2001) and
Carmen Garca and Urbina (2003), have studied correlations between the properties of crude oils and their owing properties,
including the precipitation and deposition of wax during ow.
Models have been developed to predict the onset of wax precipitation and the deposition of wax along pipeline walls. However,
accurately modelling deposition in pipelines can be a complex
and difcult undertaking, because, while precipitation is mainly a
function of thermodynamic variables such as composition,
pressure and temperature, deposition is also dependent on ow
hydrodynamics, heat and mass transfer, and solidsolid and surfacesolid interactions (Hammami et al., 2003). Only recently has
a model been developed that incorporates correct analogies for
heat and mass transfer.
This paper reviews cases where researchers have studied
ways to model wax deposition and the aging of wax deposits
in pipelines; methods of measuring wax build-up in pipelines;
methods of inhibiting this deposition; wax removal methods;
and restart procedures for pipelines gelled with waxy crude. In
doing so, this paper, as one goal, seeks to show how our understanding of these mechanisms has developed, to highlight areas
where further understanding of these mechanisms is still needed,
and to show how well our current correlations can be applied to
the accurate prediction of wax deposition. Furthermore, this paper seeks to highlight the progress that has been made in devel-

681
682
682
682
683
683
685
686
686
687
687
687
687
688
688
689
692
692

oping methods to mitigate and treat the formation of parafn


layers in pipelines.

2. Detection of deposited wax


2.1. Detecting blockages
In order to experimentally explore wax deposition in the eld or
to determine the locations of particularly large wax deposits or
even complete plugs, methods are needed for detecting the extent
of wax deposition at different points in a pipeline or of detecting
the location of plugs. Pressure echo techniques can be used to nd
the location of a blockage by measuring the time for a pressure
wave to be reected back along the pipeline from the point of
blockage (Chen et al., 2007). Alternatively, the pipeline could be
pressurized and then a special tool with a calliper and video camera on a remotely-operated submersible could be used to measure
the external diameter of the pipeline. Upstream of the blockage,
but not downstream of it, an appreciable difference in the diameter
can be detected when the pipeline is pressurized (Sarmento et al.,
2004).

2.2. Detecting wax deposits


Traditional experimental methods for measuring the extent of
wax deposits include direct methods such as pigging and the
take-out method, in which a section of pipe is removed and
the volume of wax inside measured. Additionally, pressure drop
and heat transfer methods can be used to measure wax deposits
indirectly without down time (Chen et al., 1997). Zaman et al.
(2004) explored alternative methods of measuring wax deposition
in pipelines. Firstly, they experimented with measuring light
absorption through crude oil using a light source and a detector
circuit mounted within a pipe. They found that, in laboratory tests,
this detector circuit proved capable of detecting contamination
even with a very small percentage present. The use of ultrasound
for solid detection, also explored by Zaman et al. (2004) proved
very successful in detecting extremely small solid grains. Finally,
they were able to use a strain gauge to detect very small changes
in pipeline weight associated with wax deposition. However, all
of these methods were only tested with small-scale laboratory representations of actual systems. Practical methods for application of
these tools to actual subsea pipelines would still need to be designed. Zaman et al. (2006) have also experimented with the use

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

of a laser spectroscope to detect parafn in parafn-contaminated


oil samples.

3. Wax deposition mechanisms


The behaviour of waxy crudes is usually approximated by modelling them as Bingham-like uids. Different mathematical models
have been proposed ranging from a general one-dimensional model of a waxy crude oil to models that describe crude oils depositing
wax in closed ow loops. For example, Fusi (2003) and Fasano et al.
(2004) delineate many models of differing complexity for the representation of waxy crude oils. In order to fully model the ow of
these crude oils, the mechanisms governing the deposition and removal of solid wax must be incorporated into the model. Then
models can be developed, informed by a theoretical understanding
of the mechanisms at play and the properties of the mixtures under study. However, the question arises of which mechanisms
are actually relevant.
Investigations in this area have been ongoing for decades by
researchers such as Hunt (1962); Burger et al. (1981), and Leiroz
and Azevedo (2005). Azevedo and Teixeira (2003)did a critical review of wax deposition mechanisms, starting with wax deposition
by molecular diffusion as described by Burger et al. (1981). In this
review it is acknowledged that, in most models of wax deposition,
molecular diffusion is treated as the dominant mechanism, and it is
also argued that experimental evidence suggests that gravity settling and shear dispersion play no signicant role in wax deposition. However, Azevedo and Teixeira point out that shear
dispersion may play a role in wax deposit removal, which would
affect the rate at which wax accumulates. Other authors, such as
Solaimany Nazar et al. (2005b) and Correra et al. (2007), have
incorporated wax removal mechanisms involving shear forces
(sloughing, ablation) into their wax deposition models. Other
mechanisms including thermo phoresis, the Saffman effect and
turbophoresis have also been considered in modelling wax deposition (Merino-Garcia et al., 2007).
3.1. Molecular diffusion mechanism
It is assumed that, for the ow of crude oil in the turbulent regime, the turbulent diffusivities of momentum, chemical species
and temperature will lead to a uniform distribution of velocity,
temperature and concentration proles in a pipe cross-section.
Therefore, the transport of wax will be controlled by the gradients
prevailing at the laminar sub-layer close to the wall (Azevedo and
Teixeira, 2003). In a subsea pipeline in which the walls are cooled
below the cloud point, there will be a radial temperature gradient
and wax crystallization will occur in cooler regions nearest to the
wall. Thus, solid wax crystals will exist in equilibrium with the liquid phase. Since wax solubility decreases with temperature, there
will also be a concentration gradient established by the temperature gradient within the pipeline, with the cooler regions near
the wall having the lowest concentration of wax in the liquid
phase. This is what leads to the molecular diffusion of wax from
the bulk uid to the walls of the pipeline.
Azevedo and Teixeira (2003) suggested that the mass ux of the
wax be estimated by Ficks Law as

dmm
dC
qd Dm A
dr
dt

Here mm is the mass of deposited wax, qd is the density of the solid


wax, Dm is the diffusion coefcient of liquid wax in oil, A is the surface area over which deposition occurs, C is the concentration of
wax in solution (volume fraction), and r is the radial coordinate.

673

3.2. Soret diffusion


Soret diffusion or the Soret effect refers to thermal diffusion,
which accounts for mass separation caused by the existence of a
temperature gradient within the pipeline (Ekweribe et al., 2009).
Some researchers, such as Merino-Garcia et al. (2007), have classied its effect in wax deposition as negligible. However, expressing
diffusion in terms of molecular and thermal diffusion allows for a
wax deposition model to more correctly account for thermal effects in diffusion (Banki et al., 2008). Thus, total mass ux would,
ideally, have to be represented as a combination of Ficks Law, in
terms of Dm and the concentration gradient, and transport by the
Soret effect, in terms of a thermo diffusion coefcient, DT, and
the temperature gradient.
3.3. Brownian diffusion mechanism
This would occur when wax crystals that have precipitated out
of the oil solution collide with excited oil molecules. The use of this
mechanism in modelling deposition was also explored by Azevedo
and Teixeira (2003). This diffusion mechanism can also be represented by Ficks Law as shown in equation.


dmB
dC
qd DB A
dt
dr

Here mB is the mass of wax deposited by Brownian motion, DB is the


Brownian motion diffusion coefcient of the solid wax crystals and
C is the concentration of solid wax out of solution.
Azevedo and Teixeira (2003) acknowledge that many authors
dismiss Brownian diffusion as a relevant mechanism for wax deposition. However, they conclude that there is not enough evidence
to warrant this, citing an argument used by Majeed et al. (1990),
which suggests that Brownian diffusion ux will be away from
the wall, where the solid concentration would be highest. They dismiss this argument, because if the wax crystals are trapped in the
immobile solid layer at the wall, the concentration of solid crystals
in the liquid at the wall is zero, or nearly zero, allowing for Brownian diffusion toward the wall. The review concludes that Brownian
diffusion remains a possible contributing mechanism for wax
deposition.
3.4. Gravity settling mechanism
Azevedo and Teixeira (2003) classify gravity settling as insignificant in contributing to wax deposition, citing experimental evidence from Burger et al. (1981), which showed that the settling
velocities of wax crystals under typical conditions do not contribute signicantly to deposition. This was further supported by
experimental evidence from Burger et al., which demonstrated that
deposition under horizontal and vertical ow is identical within
the limits of experimental error.
3.5. Shear dispersion mechanism
Shear dispersion could contribute to wax deposition through
the lateral motion of particles immersed in a shear ow. Some
authors, such as Fusi (2003), include deposition in terms of a shear
dispersion coefcient in the modelling of wax deposition. Also,
Fasano et al. (2004) claim that, based on the literature, for temperatures much lower than the cloud point and for moderate heat
uxes the dominant process is shear dispersion, while for slightly
higher temperatures the dominant process is molecular diffusion.
However, Azevedo and Teixeira (2003) claim that shear dispersion
does not contribute to deposition, because experimental evidence
shows no deposition of wax under conditions of zero heat ux,
when it would only be possible if driven by a ow-induced

674

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

mechanism, such as shear dispersion. However, Azevedo and


Teixeira concede that shear forces can still contribute to the
removal of wax deposits. Regardless of conicting theories with
regard to the role of shear dispersion in wax deposition, the importance of this mechanism in the overall accumulation and aging of
wax deposits cannot be ignored.
3.6. Shear stripping mechanism
Removal of wax deposits by shear forces becomes especially
important under turbulent conditions when the rate of removal
will be signicantly higher compared to laminar ow. Therefore,
in order to accurately model wax deposition, especially for turbulent ow, it is necessary to incorporate shear stripping effects into
the model. Additionally, modelling wax removal by shear forces
could help in the design of ow improver chemicals, as some of
these may act by causing the formation of softer gel structures that
are more susceptible to removal by shear forces. Some researchers,
such as Matzain (1999), have tried to represent this effect as an
empirical correlation for the reduction in rate of deposit formation
caused by shear forces.
3.7. Nucleation and gelation kinetics
The crystallization of waxes is a kinetic process, the onset of
which can be described by classical homogeneous nucleation theory (Paso, 2005). While much work has been done to approach wax
deposition as a thermodynamic problem, modelling based on the
kinetics of deposit formation has not been widely explored
(Merino-Garcia et al., 2007). Paso (2005) sought to address the
insufcient understanding of the crystallization and gelation processes, as well as the assumption that parafn precipitation kinetics does not limit deposition rates; an assumption that could lead
to the prediction of wax deposition in cases where a stable gel
cannot form. He used model uids consisting of n-parafn components dissolved in petroleum mineral oils, and applied homogenous nucleation and crystallization theory, along with differential
scanning calorimetry to measure the onset of crystallization and
the crystallization rate.
Paso (2005) compared experimental and equilibrium crystallization rates to show that there were three regimes in the crystallization process at low cooling rates. The rst is a nucleation lag
period starting at high-temperature conditions. The second is a
supersaturation growth period, driven by the supersaturation
established during the nucleation lag period as well as by decreasing solubility conditions, and during which the crystallization rate
can spike well above the equilibrium crystallization rate. The third,
meanwhile, is an equilibrium growth period, which starts when
the supersaturation ratio is diminished and the crystallization rate
converges with the equilibrium predictions of the vant Hoff relation. One thing noted by Paso about these regimes was that the
temperature span of the supersaturation growth regime was
independent of the model uid viscosity, providing evidence of
the absence of transport limitations in the crystallization rate.
Through the application of the vant Hoff solubility model within the framework of classical homogeneous nucleation theory, Paso
(2005)demonstrated that nucleation represents the primary
kinetic limitation associated with the crystallization of n-alkanes
in organic solution at low cooling rate conditions, with crystallization rate limitations becoming signicant at high cooling rates. He
also highlighted that the initial nucleation event is dependent upon
the solubility behaviour of the highest fraction of n-alkane components in the uid, and that the introduction of chain-length variations effects a reduction in the critical nucleus surface energy by
co-crystallization of dissimilar chain-length parafns.

Paso (2005) also investigated the mechanical properties of waxy


model uids at constant cooling rates using controlled-stress rheometric measurements, applying an oscillatory upon the uid samples in order to characterize their mechanical properties during
gelation. The crystal structure in samples was also studied via
microscopy and, furthermore, Paso applied an extension to an
established three-dimensional analytical percolation approximation to waxoil gel systems. This allowed for the prediction of theoretical gelation via the percolation threshold, the fractional
volume of the solid crystalline phase at which it forms a continuous, domain-spanning path connected by crystalcrystal interactions. For this purpose, parafn crystals were represented by
ellipsoidal geometries with spherical rotational volume of interaction. The primary and secondary ellipsoidal aspect ratios of the
crystals, a1 and a2, were related to the solid phase fraction at the
percolation threshold, ug, by equation.

/g hp

1 1

a1 a2

Here hp = 0.295 represents the spherical percolation threshold.


While this would give a prediction of the formation of a crystal percolation network, it was noted that this will lead to gelation only if
the number density and strength of the crystalcrystal interactions
are sufcient to impart solid-like properties to the uid (Paso,
2005).
Overall, Paso (2005) noted that the gel point of a waxy petroleum uid is dependent on the morphologies and surface characteristics of the randomly oriented parafn crystals, and that
aspect ratios on the order of 100 allowed mechanical gels to form
from these oils with parafn content as low as 0.5%. Also, that
mono disperse crystals exhibited ordered surfaces and sharp edges,
providing minimal crystalcrystal contact and weak interactions,
while polydisperse n-alkane crystals exhibited nano-scale surface
roughness, which provides contact points for strong crystalcrystal
interactions, allowing for mechanical gelation at smaller wax contents. Additionally, Paso concluded that percolation threshold
models provide accurate gel point predictions for physical gelation
systems that exhibit strong crystalcrystal interactions, while under-predicting the solid fraction necessary to induce gelation in
weakly-interacting particle systems.
Other recent studies that approached the subject of nucleation
and gelation kinetics include those by Lopes-da-Silva and Coutinho
(2007) and Ekweribe (2008). They analyzed gelation kinetics with
the phenomenological Avrami model and noted an apparent
dependence of nucleation and crystal growth mechanisms and
rates on the degree of supercooling below the WAT at which crystallization is occurring. Lopes-da-Silva and Coutinho (2007) also
noted an apparent predominance of heterogeneous nucleation
and diffusion-controlled growth, especially at higher supercooling
and/or higher oil complexity composition and molecular weight.
These results and those of Paso (2005) and further studies should
prove invaluable in the development of more robust wax deposition models, which take kinetic considerations into account. They
can also be useful in determining mechanisms by which wax gelation can be inhibited or wax deposits weakened by wax crystal
modication.
3.8. Deposition in two-phase ow
Analyzing and modelling liquidliquid two-phase ow has previously been explored by many researchers as well as present
authors (Raj et al., 2005; Chakrabarti et al., 2006, 2007). Deposition
in two-phase ow shows some characteristics similar to liquid
liquid two-phase ow. Matzain et al. (2002) found that the
thickness, hardness and prole of wax deposition in two-phase
gasoil ow show dependence on ow patterns. They used a closed

675

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

ow loop and the liquid displacementlevel detection (LDLD)


technique, proposed by Chen et al. (1997), to measure wax deposits under different conditions. For horizontal ow, the thickness of
deposits varied around the circumference of the pipe depending on
ow pattern, as shown in Fig. 1.
Matzain et al. (2002) account for these distributions by describing how, in stratied ow, only the lower part of the wall will be in
contact with the oil phase, and the heat transfer rate will be highest at the bottom of the pipe and will decrease upward, resulting in
decreasing deposit thickness in a crescent shape. In the case of
wavy stratied ow, the wavy gasoil interface is cooled because
of the waves, increasing heat transfer rate and, thus, deposit thickness along the interface. With intermittent ow, the passing of liquid slugs induces high shear force and stress along the bottom of
the test pipeline and shearing of wax deposits, resulting in thinner
deposits at the bottom of the pipe. With annular ow, the wax
thickness is uniform around the circumference, as oil is uniformly
in contact with the entire wall surface.
The results of Matzain et al. (2002) also showed changes in
hardness of the wax deposits for different ow patterns. Stratied
ow gave a soft deposit at the bottom of the pipe, with harder and
thicker deposits along the edge of the wavy gasliquid interface.
Intermittent ow resulted in a hard deposit, with increasing hardness from the top to the bottom of the pipe. Lastly, annular ow resulted in a very hard deposit, uniform across the circumference of
the pipe. Their results for vertical two-phase ow, on the other
hand, showed very uniform thickness distribution in the different
ow regimes, with very hard deposits for annular ow, deposits
of medium to high hardness for intermittent ow, and hard deposits for bubbly ow with high supercial velocity.

4. Effect of emulsied water on gelation


Crude oil emulsions, in particular, can pose signicant ow
assurance risks and, with the increase in multiphase production
in offshore environments, it has become important to evaluate
the impact of emulsied water on crude oil gelation (Visintin
et al., 2008). The presence of water over a threshold value can promote gel formation and viscous waxoil 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
as 70% (de Oliveira et al., 2010). Paso et al. (2009c) attributed the
stability of waxy emulsions to the stabilizing effect of asphaltene
particles on oilwater interfaces. They also suggested that, at
low-temperature conditions, molecular asphaltene adsorption
onto precipitated wax crystals may increase the water wettability
of the crystals, thus promoting adsorption at the oilwater
interface.
Visintin et al. (2008) hypothesized that the solid parafn stabilizes the emulsion by being strongly adsorbed at the liquidliquid
interface forming Pickering emulsions. They suggested that, by
means of the strong interaction between wax crystals and the drop
surface, growth of the gel network involves the droplets themselves, forming a volume-spanning wax crystal network with

Stratified
Smooth

Stratified
Wavy

entrapped dispersed water, as shown in Fig. 2. They observed a


sharp increase in shear viscosity, yield stress and pour point for
waxy crude oil emulsions with above 2530% volume of dispersed
water, as demonstrated in Fig. 3.
It was similarly noted by de Oliveira et al. (2010) that these viscous emulsions can increase gel strength and hinder pipeline restart by increasing the magnitude of the rheological properties of
the waxy crude oil gel. They attributed this change to the network
developed by the aggregation of the waxy crystals and water. Paso
et al. (2009c) also noted drastic increases in uid viscosities and
shear thinning rheological behaviour due to the presence of emulsied water. These observations show the signicance of considering the effect of emulsied water on gelation, and Visintin et al.
(2008) note the importance of accounting for the impact of emulsied water during eld development studies. Water fraction produced by a well generally increases over its lifetime (Lockhart and
Correra, 2005; Visintin et al., 2008). Thus it would be very useful to
account for the increasing impact of emulsied water on gelation
and gel rheology during continued operation.

5. Cloud point, pour point and gel point correlations


Some authors have focused on developing correlations between
measurable properties of crude oils, such as the pour point, and the
conditions under which disruptive wax deposition will occur.
Work such as this may help in predicting if and when fatal wax
deposition would occur in pipelines carrying particular crudes. Li
et al. (2005) cited the results of Holder and Winkler (1965) as indicating that 2 wt.% precipitated wax is sufcient 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
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 oils., represented graphically by Figs. 4 and 5. These results
and future research could be useful in both determining the tendency for different waxy crudes to gel and harden at particular
temperatures, and in devising chemical means of inhibiting this
occurrence.

6. Review of some existing wax deposition models


Many different authors have proposed models for the ow of
waxy crude oils and the associated deposition of solid wax within
pipelines, including Farina and Fasano (1997), and Fusi and Farina
(2004). Additionally, there are commercial software codes
developed to describe these processes, such as those compared
by Bagatin et al. (2008). Fasano et al. (2004) reviewed various
mathematical models for the ow of waxy crude oils in laboratory
experimental loops, in which the oils are assumed to behave like
non-Newtonian Bingham uids, a common assumption for modelling these uids. Torres and Turner (2005) approached the problem
by developing a method of straight lines for solving a Bingham
problem for modelling the ow of waxy crude oils.

Intermittent

Annular

Fig. 1. Approximation of wax thickness distribution for various horizontal ow patterns (as described in Matzain et al., 2002).

676

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

Fig. 2. Schematic representation of the gelation of waxy crude oil emulsions. Parafn 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 parafn 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).

Fig. 3. Pour point of waxy crude oil emulsion with increasing water content
(Visintin et al., 2008).

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
Hyun Su Lee.
6.1. Thermodynamic models

The earlier models presented here incorporate the wax deposition processes for pipelines containing waxy crude oils, and consider cases where either molecular diffusion or shear dispersion
is considered the dominant mechanism involved in wax deposition. However, one of the mistakes commonly introduced to wax
deposition models is the assumption that the temperature and
concentration gradients are independent, and that, therefore, wax

Many researchers have studied the thermodynamics of wax


deposition in hopes of creating a model that accurately describes
the process. In one example of earlier work, Lira-Galeana et al.
(1996)developed a thermodynamic framework for calculating
wax precipitation in petroleum mixtures as several distinct solid
phases. Solaimany Nazar et al. (2005a) later developed a

677

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

Fig. 5. Tc (2 wt%) vs. gel point (Li et al., 2005).

multi-solid phase thermodynamic model for predicting wax precipitation in petroleum mixtures, by using the PengRobinson
equation of state to evaluate the phase behaviour of both liquid
and vapour phases. The model is solved for equilibrium, in which
the fugacity of each component is equal in every phase, using Eq.
(4), proposed by Prausnitz et al. (1986).

"
!
!
 s
DHfi
f
T
DHtri
T


exp
1

1

fl i
RT
RT
T tri
T fi
#
Z Tf
Z f
i
1
1 T i DC pi

DC pi dT
dT
RT T
R T
T

Here fis is the solid phase fugacity, DHri is the enthalpy of solid
solid transition between different solid phases, T fi is the temperature of fusion, Ttr is the transition temperature, Cp is the heat capacity, and R is the ideal gas constant.
Table 1 shows a comparison of experimentally determined
WATs for ve synthetic parafn systems and those predicted by
the model of Solaimany Nazar et al. (2005a) and a UNIQUAC model
developed by Coutinho (1998). The synthetic systems were each
composed of decane and a bimodal parafn distribution. It should
be noted that with this and other models which use experimentally
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
and Daridon (2005) and Hammami et al. (2003). Therefore, agreement with experimental data may not prove denitively the accuracy of a model, especially as far as its applicability to a wide range
of waxoil mixtures.
Wuhua and Zongchang (2006) also developed a more recent
thermodynaamic model, based on the equality of fugacities at
equilibrium, which estimates solid precipitation as a function of
temperature and composition. For this study, Eq. (5) was used for
the condition of equal fugacities in the solid and liquid phases.

xSi
cL f L
iS iS exp
L
xi
ci fi

Z
0

V Li  V Si
dP
RT

and liquid phase respectively. For their model, there was an added
level of specicity for modelling particular n-alkane species. Different correlations were used for the fusion enthalpies of n-alkanes
and for their enthalpies of solidsolid transition based on both carbon 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 renement of
thermodynamic correlations, as performed by Wuhua and
Zongchang, can increase model accuracy in predicting precipitation as a function of temperature.
Further studies, for example, by Edmonds et al. (2008), have
also explored ways of representing the wax phase in order to more
accurately model wax deposition. Edmonds et al. modelled the
wax phase as a continuous distribution of n-alkane components,
showing how this eliminated physically unrealistic artefacts found
in the predictions of models that lumped n-alkanes into pseudo
components. Edmonds et al. carried out simulations with numbers
of components approaching 100 and, in order to increase the computational speed, converted phase equilibrium and physical property data into empirical expressions, tted to the rigorous model.
They also noted the importance of considering the deposit limiting
mechanism of wax shearing in order for both their model and others from the literature to more accurately agree with the limited
eld data available from actual pipelines.

6.2. Hydrodynamic model


Ramrez-Jaramillo et al. (2001) also developed a multi-solid
phase thermodynamic model for predicting wax deposition. In
addition, Ramrez-Jaramillo et al. (2004) developed a multicomponent liquid-wax hydrodynamic model for simulating wax
deposition in pipelines, which treated molecular diffusion as the
dominant mechanism. Fig. 7a shows the computational domain

Table 2
Experimental WAT data and model predictions for crude oils (Wuhua and Zongchang,
2006).
Sample

Experimental
results

Leelavanichkul
model

Crude
Oil A
Crude
Oil B
Crude
Oil C

298.2 K

298.8 K

295.2 K
294.2 K

Deviation

Present
model

Deviation

0.4 K

301.3 K

3.1 K

293.4 K

1.8 K

295.4 K

0.2 K

296.0 K

1.8 K

297.8 K

3.6 K

!
5

Here x is the mole fraction, c is the activity coefcient, V is volume, 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).

WAT (K)
UNIQUAC
This model

Bim 0

Bim 3

Bim 5

Bim 9

Bim 13

308.75
307.05
308.45

309.65
307.55
309.05

310.37
308.47
309.55

311.33
309.63
310.7

312.81
311.41
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) 671694

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

used by Ramrez-Jaramillo et al. (2004), consisting of a model pipe


of length, L, and radius, r, along which a mixture of hydrocarbons
ows. The pipe was divided to form a computational mesh, with
boundary conditions applied at the ends and along the exterior
surface of the pipe, and nite differences were used in the solution
of differential equations.
Ramrez-Jaramillo et al. (2004) modelled the uid as consisting
of n hydrocarbon components in thermodynamic equilibrium, with
mole fractions, in both the liquid and solid phases, that are functions of pressure and temperature. They considered the wax deposition rate to depend on oil composition, oil temperature, external
temperature around the pipe, ow conditions, pipeline size and
pressure. The model assumed wax deposition by molecular diffusion and removal by shear forces, which would be especially significant at high Reynolds numbers [( quDh /l), where q = density,
u = velocity, l = dynamic viscosity, Dh = hydraulic diameter]. In
addition, the model included aging by the diffusion of wax into
and within the gel-like deposit, which is discussed later in this
paper. The mass ux was calculated for all components in the
system and summed to give the total ux.

Ramrez-Jaramillo et al. (2004) used mass, momentum and


energy balances, shown in Eqs. (6)(8) and assumed mixture
incompressibility and quasi-steady state for all rate processes
concerning mass, momentum and energy.

@ qm
r  qm m 0
@t

qm


@m
m  rm rP r  s qm g
@t

qm C v



@T
m  rT kr2 T
@t

Here P, s and g are the pressure, stress tensor and gravitational


constant; Cm and k are the heat capacity and thermal conductivity
(which is assumed constant), respectively; and m is the average
macroscopic velocity of the mixture. They expressed the total
amount of deposited wax, M(t,z), in terms of the deposited mass

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

of each component due to molecular diffusion, MMDi(t,L), the mass


removed by shear forces, MSR(t,L), and the mass of wax molecules
diffusing into the gel deposit, MGD(t,L), as shown in equation.

Mt; z

n
X

M MDi t; L  M SR t; L  MGD t; L

i1

Ramrez-Jaramillo et al. (2004) solved for the total deposition


rate, @M/@t. The output of their model included solid fractions, density, viscosity, radial mass ux and deposited mass calculations.
The model showed reasonable agreement with previously developed models and experimental data for a binary mixture reported
by Cordoba and Schall (2001), as shown in Fig. 8 Ramrez-Jaramillo
et al. found that the Peclet number and Reynolds number parameters had a signicant impact on the amount of wax deposited.
7. Wax aging models
7.1. Counter diffusion
Researchers have also explored the properties of wax crystals
and wax deposits formed from crude oils. Nautiyal et al. (2008)
studied the crystal structure of n-alkane parafns crystallized from
crude oil. Other studies have specically looked at the way wax
deposits change after the initial formation. This is important because, in addition to understanding the mechanisms involved in
the deposition of wax in pipelines, in order to fully model the ow
of crude oil and accumulation of wax, it is vital to understand the
mechanisms that govern the aging of wax deposits. These deposits
are not simply static and unchanging. Rather, after a layer of wax
has formed along a pipeline wall, its composition gradually
changes. The crystalline wax deposit actually behaves like a porous
medium with oil trapped within its three-dimensional network
(Singh et al., 2000, 2001a). The wax content of this deposited gel
can therefore increase with time by diffusion. As this happens,
hardness, melting point and heat of fusion of the deposit can
change, which could affect decisions about the appropriate method
of wax removal to employ in a pipeline.
Singh et al. (2000) studied this phenomenon by use of food
grade wax dissolved in a mineral oilkerosene mixture, which
was pumped through a closed ow loop setup. Their experimental
procedure involved heating a waxoil mixture to 3035 C in a
stirred tank and maintaining the temperature of this vessel above

679

the cloud point, while pumping the waxoil mixture through the
ow loop. The ow loop consisted of a 5/8 in. OD steel tubing test
section, which was cooled by a heat exchange jacket, and an identical but non-cooled reference section. Pressure taps connected to
pressure transducers were used to measure the increase in differential pressure during operation in order to determine the thickness of the deposit within the test section. The bulk uid inlet
temperature, tb, and wall temperature, ta, were also monitored.
Singh et al. (2000) determined that a counter diffusion phenomenon, in which wax molecules diffuse into the gel deposit and oil
molecules diffuse out of the deposit, is responsible for the aging
of the deposit. They furthermore determined that the rate of aging
is dependent on oil ow rate as well as the pipeline wall temperature. In their experimental setup with oil in a closed ow loop with
cooled walls, there was a rapid decrease in internal radius measured over the rst day followed, which then plateaued. Similarly,
the increase in the measured weight fraction of wax slowed after a
rapid change in the rst day. The wax content (determined using
high-temperature gas chromatography, HTGC) of the gel deposit
also changes over time, with the proportion of lighter components
decreasing after the rst day, while the proportion of heavier components increases. The data recorded by Singh et al., showed that
the wax content of the deposit continued to increase even after
the thickness stabilized, and that waxes of chain length higher than
29 diffused into the deposit while the ones with lengths less than
29 diffused out. 29 is the critical carbon number, CCN, for the given
operation conditions; a value that could be useful in determining
what inhibitors to use in a particular well or pipeline, based on
whether or not they can inhibit crystallization of waxes above
the CCN (Paso and Fogler, 2003).
Singh et al. (2000) were able to develop a mathematical model
to describe the wax deposition process in a laboratory ow loop by
solving numerically a coupled system of differential and algebraic
equations of heat and mass transfer inside and outside the gel deposit. Eq. (10) shows the mass balance they used to relate the rate
of change of wax in the gel deposit to the radial convective ux of
wax molecules from the bulk of the uidgel interface.

d
pR2  r 2i F w tLqgel  2pr i Lk1 C wb  C ws T i 
dt

10

Here R is the original internal radius of the pipe, ri is the internal


radius during deposition (average radius available for ow of oil),
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 coefcient, 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 obtained from the energy balance shown in equation,

2pr i hi T b  T i

2pke T i  T a
 2pri k1 C wb  C ws T i DHf
lnR=r i

11

where hi is the interface heat transfer coefcient, ke is the effective


thermal conductivity of the gel, and DHf is the heat of solidication
of the wax. The heat and mass transfer coefcients 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 ow loop to the radial convective ux of wax molecules from
the bulk to the uidgel interface and the diffusive ux into the gel
at the gel interface.

2pr i F w tqgel
Fig. 8. Dimensionless wax thickness distribution vs. time. Comparison of model
predictions with experimental data for the 30:70 (cyclo C6C19:C8) ratio (RamrezJaramillo et al., 2004).



dr i
dC w 
2pr i k1 C wb  C ws T i   2pri De
dt
dr i

12

Here De is the effective diffusivity of wax inside the gel deposit.


Coupled differential equations from Eqs. (10) and (12) were solved
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) 671694

Fig. 9. Kinetic growth of crystals for oil sample from X-ray diffraction analysis at
10 C (Coutinho et al., 2003).

instant using RungeKutta algorithms, along with equations for


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. However, Lee (2008) has shown that the mass-heat transfer correlations used by Singh et al. incorrectly assume independent heat
and mass transfer, and were successfully applied only because of
a high degree of supersaturation in the laminar boundary layer.
Other researchers, such as Hernandez et al. (2004), modelling
wax deposition in pipelines has begun incorporating wax aging
into their models. Additionally, Singh et al. (2001b) were able to
develop a thermodynamic model to predict both cloud point temperatures and CCNs of waxoil mixtures, where CCN is a function
of the mixture composition as well as the wall temperature. This
model also showed good agreement with experimental data, predicting the cloud points and the CCNs of model oils with good
accuracy.
7.2. Ostwald ripening
It must be noted that the diffusion mechanism used by Singh
et al. (2000, 2001a,b) is not the only possible mechanism for
explaining the aging process. In fact Continuo et al. (2003) found
that aging of wax deposits takes place even for samples kept under
isothermal conditions. The diffusion mechanism for aging cannot
account for this as it is driven by temperature-composition gradients. Coutinho et al. reported broadening of peaks on X-ray diffraction and Cross Polar Microscopy (CPM) images showing an increase
in the crystallites size. Fig. 9 shows an example of their results
from X-ray diffraction analysis, for which the crystallite size, r, is
related to a shape factor K, and the measured peak position, h,
and breadth, b, by equation.

Kk
b cos h

13

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. 10. They noted that this can only occur when the heat of crystallization released is used by the melting of an equivalent mass of
crystals. This indicates that wax deposits in crudes suffer recrystallization. Coutinho et al., thus, conclude that Ostwald Ripening is
also a mechanism responsible for the aging of wax deposits.

8. Correct analogies for correlated heat and mass transfer in


turbulent ow
Many existing wax deposition models assume that heat and
mass transfer can be related by the chain rule, which assumes that
the system is at thermodynamic equilibrium (which may not be
true), or use mass-heat transfer analogies, such as the Chilton
Colburn analogy, which are valid only when the temperature and
concentration elds are independent. Venkatesan and Fogler
(2004) noted that such heat-mass transfer analogies are not applicable for predicting the mass transfer rates in turbulent ows,
where the concentration eld is correlated to the temperature eld
and the concentration boundary layer and temperature boundary
layer thicknesses are not independent. They showed that use of
the Colburn analogy in this case would result in a signicant
over-prediction of wax deposition. They also proposed a method
for estimating the convective mass transfer rate using the Nusselt
number and the experimentally obtained solubility curve. However, this method would only be valid for thermodynamic equilibrium in the mass transfer boundary layer, when precipitation
kinetics are not limiting.
For the development of more rigorous and accurate models, it
has been necessary for researchers to explore the correct relationship between heat and mass transfer. Lee (2008) investigated the
combined heat and mass transfer phenomenon under laminar
and turbulent ow conditions using the nite difference method.
He developed a model based on that of Singh et al. (2000), which
could be applied for any precipitation kinetics. For turbulent ow,
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) 671694

and Fogler (2004) under-predicts deposition by assuming that the


concentration prole in the mass transfer boundary layer follows
the thermodynamic equilibrium limit between temperature and
concentration at every point. This was contrasted with the overprediction of the ChiltonColburn analogy, which gave maximum
supersaturation. The comparison showed that these two approaches constitute the limiting cases for deposition, and that
the actual concentration prole, which is dependent on the precipitation kinetics, falls between those calculated by the two methods.
Instead of using those two limiting cases, Lee (2008) employed
a computational approach for calculating the Nusselt numbers
[(hL/kf), where L = characteristic length, kf = thermal conductivity
of the uid, h = convective heat transfer coefcient] and Sherwood
numbers [(KL/D) where L is a characteristic length, D is mass diffusivity, K is the mass transfer coefcient] according to following
equations.

Nu

Sh



2ri @T
@r rr
Tb  Ti


2r i @C
@r rr
Cb  Ci

2r i hi
k

14

2r i kM
Dw0

15

The temperature and concentration gradients at the uid-deposit interface, needed for these calculations, were obtained by
solving mass and energy balance equations, as shown in following
equations.

mz



@C 1 @
@C

rDwo
 kr C  C ws
@z r @r
@r

16

mz



@T 1 @
@T
 bC  C ws

r aT
@z r @r
@r

17

Here vz is the axial velocity, Dwo is the molecular diffusivity of


wax in oil, kr is the thermal conductivity, aT is the thermal diffusivity, and the precipitation term b(CCws) is considered negligible.
Lee rst did this for laminar ow. Using a discretized form of the
mass-heat transfer equation along with their appropriate boundary conditions, Lee wrote the governing equations in matrix form.
Then by inverting these matrices to give the radial temperature
and concentration proles, and numerically marching from the inlet of the tube to the exit he could obtain the complete set of temperature and concentration proles with respect to the radial and
axial position.
From this, Lee (2008) showed how the Sherwood number
prole as a function of axial distance would change for different
precipitation rate constants. This showed that if there was no precipitation in the boundary layer, the heat and mass transfer rates
become independent of each other, resulting in a supersaturation
curve. However, as the precipitation rate constant increases the
Sherwood number is decreased, because wax molecules would
not reach the oildeposit interface, and would instead ow down
to exit as solid particles.
To obtain the Sherwood and Nusselt numbers under turbulent
conditions, Lee (2008) used the same procedure with governing
equations modied for turbulent ow to include the turbulent
axial velocity prole and the thermal and mass transfer eddy diffusivities. The wax concentration proles in the turbulent boundary
layer obtained this way showed that heat and mass transfer
become independent as the precipitation rate constant approaches
zero, resulting in the ChiltonColburn analogy-derived concentration prole. Conversely, as the precipitation rate constant
increases, precipitation in the boundary layer increases, with concentration approaching the solubility limit for thermodynamic
equilibrium.

681

In his model, after calculating the Sherwood and Nusselt numbers, Lee (2008)could then solve the growth and aging governing
equations from Singh et al. (2000)s model to solve for deposit
thickness and wax fraction at each time step in his computational
procedure. Lee (2008)showed that there was excellent agreement
between the results of his model and lab-scale laminar ow loop
experimental data. There was also good agreement with turbulent
lab-scale results, though there was signicant discrepancy for early
times at higher volumetric ow rates, possibly due to sloughing.
The results of the computational model also closely matched
large-scale ow loop data. The results obtained by Lee (2008) show
that this model is applicable for varying precipitation kinetics, and
provides a robust and rigorous way of predicting wax deposition
under a range of turbulent conditions.
9. Inhibition of wax deposition
The most effective way of dealing with the problem of wax
deposition in crude oil pipelines would be to prevent it from occurring in the rst place. Thus, researchers have investigated different
methods of inhibiting the deposition process. These include the
heat insulation of subsea pipelines to actually inhibit precipitation
by keeping pipeline temperatures as high as possible (Quenelle and
Gunaltun, 1987), the internal coating of pipelines with plastics
(Patton, 1970; Bummer, 1971), and also methods of preventing
wax deposition on pipeline walls, such as the use of chemical
inhibitors, which will be discussed in more detail in this paper.
9.1. Chemical inhibitors
Many researchers have studied the efcacy of different inhibitors of wax deposition and the mechanisms by which they inhibit
this deposition, including Jorda (1966), Mendell and Jessen (1970),
Fulford (1975), Addison (1984), Newberry and Barker (1985),
Fielder and Johnson (1986), Singhal et al. (1991), Jang et al.
(2007), and Tinsley et al. (2007). The efcacy of commercially
available inhibitors tends to be limited, and has to be evaluated
on a case-by-case basis. Wang et al. (2003), for instance, found,
when testing some wax inhibitors, that the inhibitors they had
studied reduced the total amount of deposition, but had only limited success in suppressing the deposition of the high molecular
weight parafn components (C35 and above). This resulted in harder wax deposits than in the absence of an inhibitor. They also found
that inhibitors most able to depress the WAT were more likely to
be superior products for decreasing total wax deposition, and that
the addition of the corrosion inhibitor, oleic imidazaline (OI), signicantly increased the efcacy of deposition inhibition. Fig. 11
shows some of their results, where PIE is the parafn inhibition
efciency, the amount of wax deposited with inhibitor as a wt.%
of amount deposited without it.
Bello et al. (2006) also studied the efcacy of commercial wax
inhibitors, particularly on Nigerian crude oils. They found that
the use of a trichloroethylenexylene, TEX, binary system as an
additive was actually more effective and economically feasible
than the use of commercial inhibitors. Other researchers have
noted the need to tailor inhibitor treatments to particular crudes
in order to maximize efcacy. Manka and Ziegler (2001), for instance, found that additives work best when matched to the parafn distribution in the crude oil being treated. Similarly, Carmen
Garca (2001) noted a strong relationship between a specic parafn inhibitors efciency and the crude oil composition, which
would require case-by-case consideration for selecting inhibitors
for use in the eld.
Additionally, there is the consideration of the environmental
conditions under which a wax inhibitor is to be used, since, for

682

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

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

operations at particularly low temperatures, the inhibitor formulation must be winterized to allow effective delivery under those
conditions (Manka et al., 1999; Jennings and Breitigam, 2009).
Also, while work continues towards developing new, more effective wax inhibitors, it remains the case that inhibitors typically
do not provide 100% inhibition, and so are used in conjunction with
remediation methods such as pigging (Jennings and Breitigam,
2009; Kelland, 2009).
9.2. Types of chemical inhibitors
There are different mechanisms by which chemical inhibitors
can prevent wax deposition or gelling in pipelines. They can lower
the WAT or pour point or can modify the wax crystals so as to prevent their agglomeration and deposition (Kelland, 2009). The
chemicals that modify the WAT are usually referred to as wax
inhibitors or wax crystal modiers, while those that affect the pour
point are known as pour point depressants (PPDs) or ow improvers; although there is a great deal of overlap in terms of the chemistry and mechanisms of these two classes (Kelland, 2009). Some
detergents or dispersants that act as wax inhibitors, such as polyesters and amine ethoxylates, may act partly by modifying the surface of the pipe wall, rather than just the wax crystals, to prevent
adhesion (Pedersen and Rnningsen, 2003), and many effective
wax inhibitors create weaker deposits that are more easily removed by shear forces (Manka et al., 1999; Kelland, 2009). The
main types of wax inhibitors and PPDs include ethylene polymers
and copolymers, comb polymers and assorted other branched

polymers with long alkyl groups, such as alkyl phenolformaldehyde, which are not as effective as comb polymers when acting
on their own as ow improvers (Kelland, 2009).
9.2.1. Ethylene copolymers
This group includes ethylene/small alkene copolymers, ethylene/vinyl acetate (EVA) copolymers, and ethylene/acrylonitrile
copolymers (Kelland, 2009). More specically, examples of these
polymers used in wax inhibition studies include poly (ethyleneb-propylene) and poly(ethylene butene) polymers (Tinsley et al.,
2007; Kelland, 2009). Random, low molecular weight EVA copolymers, illustrated in Fig. 12, are widely used and investigated as wax
inhibitors (Kelland, 2009). The effectiveness of the EVA copolymer
as an inhibitor is inuenced greatly by the percentage of vinyl acetate in the copolymer. The, more polar, vinyl acetate content aids
solubility and lowers crystallinity and so is necessary for the
depression of the WAT, whereas the polyethylene content is necessary to allow for co-crystallization with structurally similar wax,
but, on its own, has little effect on crystallization (Kelland, 2009).
9.2.2. Comb polymers
Comb-shaped polymers, illustrated in Fig. 13, have been studied
extensively as wax inhibitors by researchers such as Duffy and
Rodger (2002), Duffy et al. (2004), Jang et al. (2007), and Soni et
al. (2008). They are usually made from (meth)acrylic acid or maleic
anhydride monomers, or both, and generally provide improved
wax inhibition compared to the ethylene copolymers (Kelland,
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) 671694

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


into a gel structure by introducing defects or repulsive forces (Jang
et al., 2007; Soni et al., 2008; Kelland, 2009). As illustrated by Figs.
14 and 15, they can accomplish this by providing nucleating sites
for wax crystals on their parafn-like pendant chains while a polar
backbone impedes the formation of an interlocking wax network
(Soni et al., 2008).
In selecting the most effective comb polymers for use with a
particular crude oil, researchers have found that the length of the
side chains plays an important role. For example, Manka and
Ziegler (2001) found that matching the average pendant chain
length of comb polymer PPDs with the parafn distribution of a
crude oil provided the greatest pour point depression. Also, Jang
et al. (2007) obtained results which suggested that using comb
structures with side arms of such length as to interact favourably
with the fraction of oil most likely to crystallize into the hard
wax phase provided the best wax inhibition. This creates a problem for especially long-chained waxes, for which it would be difcult to introduce a comb polymer (or ethylene copolymer) of
sufcient length to provide efcient inhibition, and also makes it
important to have a range of comb polymers available for treatment of different crudes (Kelland, 2009).
9.2.3. Wax dispersants
These are surfactants that adsorb onto pipe surfaces and reduce
the adhesion of waxes to those surfaces, possibly by changing the
wettability of the pipe surface to water-wet, or by creating a weak

layer from which wax crystals are easily sheared off, or by adsorbing onto the wax crystals and reducing their tendency to stick
together (Kelland, 2009). Some researchers have worked on developing their own dispersant formulations. Groffe et al. (2001), for
instance, developed their own inhibitor that shows wax dispersant
behaviour and anti-sticking properties. They suggest that this
chemical, referred to as P5, interferes with the wax crystal growth
mechanism by preventing the formation of a three-dimensional
network, and thus reduces the pour point and improves the ow
characteristics of crude oils. Fig. 16 shows the effectiveness of P5
in preventing the adherence of wax from crude oils to a steel
surface.
Typical, low-cost wax dispersants include alkyl sulfonates, alkyl
aryl sulfonates, fatty amine ethoxylates and other alkoxylated
products, but these dispersants have shown limited effectiveness
in the eld when not blended with polymeric wax inhibitors (Kelland, 2009). Dispersants, however, have been used successfully to
support the functions of polymeric ow improvers because of their
ability to hinder wax settling and deposition (Al-Sabagh et al.,
2007).
9.2.4. Polar crude fractions
It has been found that polar extracts from crude and distillate
oils, which can be extracted using super critical gases such as carbon dioxide or ethylene, and which contain asphaltenes, resins and
aromatics, can be a potential source of low-cost ow improvers
(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) 671694

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 parafn gels in crude oil. They


found that the addition of asphaltenes depressed the gelation temperature of model waxoil mixtures, as summarized in Figs. 17 and
18; although they found that beyond certain thresholds, further

addition resulted in macroscopic phase separation of the mixture,


attributable to gravity settling.
Kriz and Andersen (2005) also studied the effect of asphaltenes
on wax crystallization in crude oils. They found that this effect

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

685

certain degree. These are all properties associated with PPDs and
proposed mechanisms for their action, which include co-precipitating 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 ow improvers such as comb polymers.

Fig. 17. Gelation temperature depression (cooling rate of 1 C min1) for a foodgrade parafn wax (Wax 1) and a laboratory-grade parafn wax (Wax 2) by
addition of asphaltene (Venkatesan et al., 2003).

Fig. 18. Depression in yield stress of Wax 1 system (at temperature, Tys, below the
gelation temperature) upon asphaltene addition (Venkatesan et al., 2003).

depends strongly on the degree of asphaltene dispersion or occulation more than on the asphaltene type or origin. They reasoned
that the asphaltenes, when well dispersed at very low concentrations, are easily accessible for any kind of interaction with the parafns 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 asphalteneparafn 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 molecules occulate together when over a critical concentration, with
possible co-precipitation with waxes, resulting in an unorganized
asphalteneparafn 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 another study by Yang and Kilpatrick (2005) indicated that asphaltenes and waxes do not co-precipitate in solid organic deposits.
In accounting for the observed ow improver properties of
asphaltenes, Venkatesan et al. (2003) noted that asphaltenes have
polar groups as well as alkane chains and are soluble in oil up to a

9.2.5. Short-chain alkanes


Senra et al. (2008) analysed how n-alkanes impact the crystallization 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
et al. (2009) indicate that the ability of a particular short-chained
n-alkane to inhibit gel formation by a longer-chained one depends
on the particular pairing. The trend of this inhibition was found
to depend on the extent of differences in size and solubility characteristics 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 longchained n-alkane, polydispersity and co-crystallization weaken the
gel formed in spite of the fact that more crystallizable wax is present 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 accommodate the C32 crystals that co-crystallize with the C36. This would
make the formation of large crystals and a volume-spanning
network gel more difcult, in the same way that the inclusion of
polymeric ow improvers into wax crystal structures inhibits
aggregation and gel formation. The addition of increasing concentrations of the shorter-chained co-crystallizing n-alkane, however,
resulted in a minimum pour and gel point followed by an increase.
This was accounted for by a limit to how much the addition of the
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
co-crystallize, such as in a C36/C28 system, Senra et al. (2009) saw a
very different trend. In this case, low concentrations of the shorterchained n-alkane had no effect on the pour and gel points, until a

Fig. 19. Effect of varying the wax percent of C28 and C32 on the pour points and gel
points of 4% C36 solutions in dodecane (Senra et al., 2009).

686

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

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


was seemingly too small to inuence 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
waxoil gel formation can help signicantly in implementing
inhibition measures.
9.3. Surfaces that prevent wax deposition

Fig. 20. Effect of varying the wax percent of C28 and C30 on the pour points and gel
points of 4% C32solutions in dodecane (Senra et al., 2009).

concentration at which a sharp decrease was witnessed in both


followed by a gradual increase. This was accounted for by the fact
that, at low concentrations, the more soluble shorter-chained nalkane will not crystallize out and will not be present in high
enough concentration to disrupt the crystallization of the longerchained n-alkane, so will have no effect. Then, at a high enough
concentration, the association of the shorter-chained n-alkane
molecules with the longer-chained n-alkane crystals would disrupt
gel formation. Then gelation will occur as the more soluble shorterchained alkane is added in high enough concentration to crystallize
sufciently to form a gel. Senra et al. supported this analysis with
the results of cross-polarized microscopy experiments.
Furthermore, for a C32/C30 system, Senra et al. (2009) noted that,
due to the very similar chain length and solubility characteristics,
there was only a slight initial decrease in pour point due to the formation of co-crystals, which would have relatively few vulnerable
points since the two n-alkanes are so similar. Beyond that, the C32/
C30 system behaved much like a monodisperse system. Also, for a

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 operation. With a proper understanding of the mechanisms by which
waxes adhere to oil pipeline walls it would be possible to create
pipelines in which the nature of the walls makes adhesion unfavourable. Paso et al. (2009a) performed a comprehensive review
of the use of non-stick and anti-adhesive coatings for inhibiting solidliquid deposition phenomena, including the use of metal surface treatments and synthesized polymers. The classes of
materials that they found promising included uoro-siloxanes, uoro-urethanes, oxazolane-based polymers and hybrid diamondlike carbon and polymer coatings.
Fig. 21 shows some of the reported surface free energies of surfaces for parafn control investigated by Paso et al. (2009a), which
gives an indication of the ability for waxes to interact with those
surfaces, and thus the potential of these surfaces for preventing
wax deposition. Further study of the mechanisms involved in
wax adhesion, hopefully, will result in even more effective surface
treatments in the future.
9.4. Cold ow
Heating or insulation of subsea pipelines can be used to try to
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) 671694

very different method of inhibiting wax deposition on pipeline


walls, discussed by Merino-Garcia and Correra (2008), is the use
of Cold Flow technology. This approach suggests that it might be
possible to prevent deposition on pipeline walls by reducing the
bulk temperature within the pipeline to be equal to the temperature of the sea water around it, thus eliminating the temperature
gradient. This would allow for the waxes to be transported as a solid dispersion within the bulk uid. While wax deposition may, in
fact, become negligible in the case of zero heat ux, even below the
WAT, much work would still be required to develop the technology
required for effectively cooling the bulk uid to this condition and
for transporting the resulting cold slurry over long distances. Ilahi
(2005) also discussed SINTEF and NTNU Cold Flow technology.
Additionally Haghighi et al. (2007) and Azarinezhad et al. (2010)
proposed a wet cold ow-based concept, termed HYDRAFLOW,
for preventing gas hydrate agglomeration, with the potential
benet of wax inhibition. Gas hydrates are the solid solutions of
gas components and water. Hammerschmidt (1934) discovered
the formation of hydrates in natural gas systems. Hydrates like
waxes have concerned deep-water production at seaoor depths
of 13 km and temperatures between 2 and 4 C (Gudmundsson,
2002), conditions which encourage hydrate plug formation. Several
studies have been done on kinetics of hydrate formation. Those
previous studies can be categorized into two main subjects:
nucleation and growth. In contrast to previous studies, gas pipelines hydrate agglomeration plays an important role. After the
break-up of the hydrate lm along the interface, hydrate particles
agglomerate to form a hydrate plug (Lingelem et al., 1993) like
wax. Herri et al., 1999 analyzed the particle size distribution of
hydrate particles with the particle balance equations and a mass
transfer model. However it is difcult to describe agglomeration
from experimental observation. The particles start to agglomerate
just after the nucleation process (Mersmann, 2002). The observed
particle size distribution is a result of kinetic contributions such
as nucleation, growth, agglomeration, breakage, and attrition.
Viscosity is also a contributory factor to particle agglomeration
(Mersmann, 2002).
10. Wax removal methods
If wax deposition cannot be prevented, then it is imperative to
regularly remove accumulated wax from the inside of pipeline
walls in order to prevent the total blockage of the line. Several
methods have thus been developed for the removal of wax deposits, including complete blockages of pipelines. Traditional methods
of wax removal in the petroleum industry have always had problems and limitations, and they include mechanical removal, the
use of bottom hole heaters, the use of exothermic reactions such
as that between magnesium bars and hydrochloric acid, and the
use of parafn solvents (Woo et al., 1984). Research continues to
be done to nd the most efcient, cost-effective and safe methods
of removing wax deposits and blockages. Furthermore, some
researchers have worked on modelling the operating conditions
necessary for the successful and safe restart of gelled pipelines,
in which gelled waxy crude needs to be displaced using applied
pressure.
10.1. Pigging
The practice of pigging is a way in which wax removal is commonly accomplished in the eld. With this method, deposited wax
is mechanically removed by launching a pipeline pig along the line
to scrape wax from the walls as it is forced along by the oil pressure. This, however, poses the risk of forming a wax plug downstream from the pig as the scraped wax accumulates and is

687

compressed ahead of the pig. In such an event the pipeline could


be lost. The use of bypass pigs tries to address this problem. When
the differential pressure across such a pig becomes too high, because of the accumulation of solid wax and debris ahead of it,
the bypass pig allows liquid to ow through it and disperse the
accumulated solid ahead. However, there is always the danger that
if pigging has to be temporarily suspended due to mechanical failure, or that if the pigging frequency for a pipeline is not correctly
optimized, that the result will be a stuck pig and sizable production
losses (Fung et al., 2006).
Wang et al. (2008) studied the use of regular and bypass pigs in
the removal of wax from pipelines in a laboratory system. The test
facility used consisted of a 20 ft test section of carbon steel pipe, a
mineral oil tank, a pump to push the pig with liquid as in real pigging operations, and a receiving tank to observe the structure of
the pigged materials. Four pressure transducers were installed to
monitor pressure change along the test section during pigging
operation. Candle wax with different oil contents was cast as a lm
or plug for measuring wax breaking force or plug transportation
force, respectively. After casting, the waxy spool pieces were
mounted on the test section and the pig was pushed through the
pipe by oil from the pump, removing the wax lm or plug while
the pressures at four locations along the test section were
recorded.
They concluded that the wax breaking force increases with the
decrease in oil content and the increase in wax layer thickness;
transportation force per unit plug length is affected by oil content;
transportation force decreases with the presence of oil due to
lubrication effects; and bypass pigs exhibit a very similar breaking
force behaviour when compared with regular pigs. Other studies
have focused on determining the optimal frequency of pigging to
maintain a pipeline and avoid plug formation.
10.2. Inductive heating
Another possible wax removal process, studied by Sarmento et
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
when the pipeline is completely blocked in a horizontal section so
that it is impossible to ow chemicals to the blockage. They tested
this method using the experimental setup shown in Fig. 22. They
found that the steel layers which compose commercial exible
lines can be heated by induction and the heat transferred to a solid
wax plug in the interior of the line. They also found that their
mathematical model, which agreed well with available experimental results, suggested that the power levels required for large-scale
inductive heating might be feasible for removing wax blockage in
eld applications with undersea pipelines.
10.3. Biological treatment
Biological wax removal methods have also been studied in recent years by researchers such as Rana et al. (2010), who developed
systems of parafn-degrading bacterial consortiums with nutrient
supplements and growth enhancers for controlling parafn deposition in the tubular and well bore region and in surface ow lines.
Their results showed that their systems were highly effective,
eliminating the need for repeated scrapings of wax over a period
of several months. These methods are especially important because, if successfully implemented, they have the benet of providing continuous control of wax deposition in pipelines through
constant biodegradation, rather than just providing a very temporary x.
Etoumi et al. (2008) studied the use of Pseudomonas bacteria for
the reduction of wax precipitation in waxy crude oils. Their results

688

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

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


hydrocarbons such as kerosene, toluene, xylene and crude oil, an
effect also studied by others, such as Sifour et al. (2007). The observed overall effect of Pseudomonas treatment on crude oil
showed a reduction in the concentration of long-chain hydrocarbons (C22+). Etoumi et al. concluded that Pseudomonas species
may be an efcient species for reducing parafn deposition, and
that the speed of the biochemical action on crude oil is faster within the rst 7 days. They also concluded that an observed reduction
in viscosity and WAT is indicative of the conversion of long-chain
alkenes to short ones.
Additionally, He et al. (2003) determined through eld tests,
that two Bacillus species and a Pseudomonas species showed good
parafn removal properties in test wells, increasing oil production
and eliminating the need for more expensive wax removal processes. Thus, biological wax removal methods may prove to be
quite effective and economically benecial and warrant further
study. If a biological system can be successfully and cheaply applied under the conditions in subsea pipelines then it will provide
an extremely effective method of controlling wax deposition.
11. Restart of gelled pipelines
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
the formation of a gel throughout the pipeline consisting of precipitated wax in a viscous matrix (Chang et al., 1999). This occurrence
complicates the restart procedure, as the gelled oil would need to
be displaced in order to resume normal operations. Numerous
researchers have addressed this problem, including Smith and
Ramsden (1978), Chang et al. (1999), Davidson et al. (2004),
Frigaard et al. (2007), and Vinay et al. (2007). Chang et al. (1999)
modelled the isothermal restart of gelled pipelines by the application of higher than normal operating pressures. In this start-up
scenario, oil is pumped into the gelled line at high enough sustained pressure to overcome the static yield stress of the gel, thus
breaking up the blockage and clearing the line.
The viscoplastic nature of waxy crude oils and their timedependent behaviour complicate modelling. In order to describe
the breakdown of the gel structure along with a decrease in viscosity, Chang et al. (1999) dened the static yield stress of the gel, ss,

as the critical shear stress value for determining whether the start
of a ow from a rest state will occur. Furthermore, they dened the
dynamic yield stress, sd, as the parameter for describing the relationship between shear stress and shear rate in a ow state after
yielding. However, the description of this yielding behaviour has
seen many variations and disagreements among different authors.
Many studies have been published regarding the rheology of
waxy crudes and their gels, the dependence of gel properties on
shear and thermal histories, and how they yield (Wardhaugh
et al., 1988; Chang et al., 1998, 2000; Lopes-da-Silva and Coutinho,
2007; Lee et al., 2008; Oh et al., 2009). Wardhaugh and Boger
(1991), for instance, dened yield stress as the shear stress at
which the gelled oil ceases to behave as a Hookean solid, and referred to bulk yielding phenomena, when gross yielding behaviour
is observed, as the yielding stress or yielding point. Houwink
(1958) described a transition from elastic behaviour to plastic
behaviour and then to viscous ow, distinguished by a lower and
a higher yield stress. Meanwhile some researchers, such as Barnes
(1999), who noted the high degree of variation in the denition of
yield stress, maintained that no real yield stress exists, even for
very non-Newtonian liquids. They argue this because these liquids
continue to ow or creep even below an apparent yield stress.
Barnes notes, however, that the concept of a yield stress is useful
for describing behaviour over a limited range.
11.1. Time-dependent gel degradation
Time-dependent gel-degradation is one of the important complications in modelling the restart of gelled pipelines. Ongoing efforts to model the time-dependent rheology of gels, in order to be
able to model gel breakdown under stress, draw on research such
as that of Cheng and Evans (1965), Petrellis and Flumerfelt (1973),
and Rao et al. (1985). Recent studies of the time-dependent rheological behaviour and breakdown of waxoil gels include that done
by Paso et al. (2009b), in which the mechanical behaviour of a
model waxoil gel was examined under various shear rates. Paso
et al. observed a convergence of shear stress values for different
shear rates at absolute strain magnitudes greater than 0.1, as
shown in Fig. 23. This was indicative of gel strength following a
path-independent function of the absolute strain imposed on the
gel, and further indicated that the gel structure is a point function
of the absolute strain. Based on this they concluded that, in

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

689

their model waxoil gel at constant shear rate conditions. They


concluded that a well-dened mechanism controls the rupture of
wax crystalcrystal 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 eld with a variety of hydrocarbon 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 stressstrain data
obtained while increasing the shear rate.
11.2. Examples of restart models
Fig. 23. Measured shear stress during breakage of a waxoil gel at shear rates
ranging from 105 s1 to 1 s1 (Paso et al., 2009b).

modelling the breakdown of a wax gel at low shear rates, the entire
shear history could be represented by a single dimensionless variable in the form of the absolute strain.
Paso et al. (2009b), furthermore, determined that the maximum
shear stress did not provide a useful parameter to characterize the
gel structure. Thus, in order to dene a structural parameter, k, representing the fraction of unbroken crystalcrystal linkages remaining in the gel structure at a given shear stress, they did so in terms
of the experimental stress near the convergence point. They then
used this structural parameter in an nth order degradation model
to describe the gel breakage model, as shown in following
equation.

1
1
k  ke 1n 
k0  ke 1n ac_ b t
1n
1n

18

Here k0 and ke are the initial and equilibrium structural parameter


values, and the degradation rate parameters, n and ac_ b , were determined by tting experimental values of k to equation (18) via a least
squares minimization procedure, with ke assumed to be 2  103.
Paso et al. (2009b) were able to obtain good model ts to experimental values, as shown in Fig. 24. Their tted degradation order
for different shear rates ranged from 2.7 to 3.33, indicating that a
third order degradation mechanism controls the breakdown of

Fig. 24. Comparison of experimental and tted k values at a shear rate of 103 s1.
The optimized reaction rate order is 3.07, with a rate constant of 0.131 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
by Houwink (1958). The three yield stress model utilized an elastic-limit yield stress, se, described as denoting the materials limit
of reversibility; a static yield stress, ss, described as the minimum
shear stress required to cause the deformation of a material that
may be described as yielding; and a dynamic yield stress, sd, described as the shear stress at zero shear rate, extrapolated from
the ow curve. Chang et al. used this model to describe the three
possible outcomes of applying constant pressure to a gelled pipeline in terms of the relationship between the wall shear stress,
sw, applied to the pipeline and the initial gel strength of the oil:
 Start-up without delay (sw > ss) Flow begins immediately with
three different regions, as shown in Fig. 25, where R is the total
radius of the pipeline and rf and rc denote the boundaries of the
regions:
Flow area The outermost region (R > r > rf), consisting of a
sheared annulus. Local stress is higher than the static yield
stress (s(r) > ss). The gel structure in this region is immediately broken down and the oil becomes liquid-like, displaying a dynamic yield stress.
Creep area Middle region (rf > r > rc). Local stress is lower
than static yield stress, but higher than elastic-limit yield
stress (ss > s > se). Gel structure in this region begins to
degrade with a viscoelastic deformation.
Elastic deformation area Innermost region (r < rc). Local
stress is lower than elastic-limit yield stress (s < se). Solidlike 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 deformation area exist and, initially, no ow occurs. Flow only begins
once gel in the creep region has sufciently 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 unaffected 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 because of differences in the rate of structural breakdown caused by
differences in local shear stress. Therefore, their model takes into
account the time-dependent rheology of the waxy crude oils. A
time-dependent Bingham-style equation, shown in Eqs. (19a)(19c) was used for an approximation of the time-dependent,

690

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

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 controlled stress conditions.

s sy t gtc_ ; s > sy t
sy t

sy 0  sy 1
1 kt

sy 1

gt constant

19a
19b
19c

Here g is the plastic viscosity, c_ is the shear rate, sy is the apparent yield stress governing the behaviour of the oil, and k is a rate
constant. sy(0) and sy(1) are the apparent yield stress at times
t = 0 and t = 1, respectively. sy(0) would be equivalent to ss(0),
with an initial wall stress above this value resulting in an instantaneous nite ow rate, and sy(1) coincides with se(0), with wall
stresses below this value resulting in reversible deformation and
no possible ow.
The basic physical model used by Chang et al. (1999) to describe
the start-up process was the pumping of an incoming uid (ICF)
into a pipe of length, L, and inside diameter, D, to displace the outgoing uid (OGF), as shown in Fig. 26. Here Z(t) is the length of the
pipe occupied by the ICF, r I t and r o t are the unsheared plug radii of the ICF and OGF respectively at time, t, P1 is the inlet pressure, P2 is the exit pressure, and Pz is the interface pressure. The
radius of the unsheared plug in the ow was given by equation,

r  t R

sy t
swo t

20

where swo(t) is the wall shear stress in the OGF at time t.

For the case of start-up without delay (or start-up with delay at
time, t > tdelay), Chang et al. (1999) dene the initial wall shear
stress in terms of the pressure drop. To model the time- and position-dependent changes in the ow properties of the OGF, Chang
et al. (1999) used a nite differences method. M time intervals
were used to divide the duration of the ow from start-up
(Dt = titi1), and the ow was treated as approximately steady
in each time interval for sufciently small Dt. The sheared annulus
(r < r R) was divided, for each instant, ti, into N radial elements of
thickness, Dr, and distance, rj, from the centre of the pipe
(rj = rj1 + Dr = r + jDr). The volumetric ow rate, Qi, at time ti
was thus given by following equation,

Qi

N
X

Q j Q plug

21

j1

where Qplug is the ow rate of the unsheared plug and Qj is the


volumetric ow rate of the jth annular element. In a later work,
However, Davidson et al. (2004) maintained that the nite differences method was unnecessary, because of the quasi-steady state
assumption for the OGF, which was represented as a Bingham uid
that would have apparent yield stress and plastic viscosity independent of the shear rate and thus the radial position.
Using their model, Chang et al. (1999) could calculate the plug radius, r, for each time interval using equations (19b) and (20) with a
known wall stress. The ow rate could then be computed from j = N
at the pipe wall inward to the unsheared plug at j = 0. The onset of
turbulence was predicted by calculation of a critical Reynolds number, with the appropriate adjustment to the friction factor used in
the model. For these calculations the pipe dimensions (L and R),

Fig. 26. Schematic diagram of two-uid displacement model. (a) True interface; (b) simplied interface (Chang et al., 1999).

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

pump pressure (DPc), properties of the OGF (sy(0), sy(1), g(t), k and
qo), and properties of the ICF (sB, gB and qI) need to be known. Here qI
and qo are the uid densities of the ICF and OGF; sB is the Bingham
yield stress; and gB is the Bingham plastic viscosity. Therefore, the
accuracy of this model in predicting the time-dependent ow properties during start-up and the time needed to clear a blockage depended greatly on comprehensive knowledge of the system,
requiring accurate experimental measurements.
Davidson et al. (2004) developed another model for the restart of
gelled pipelines. This model extended the one developed by Chang
et al. (1999) to account for the compressibility and inhomogeneity
of the gelled oil and displacing uid. In this model, when the inlet
pressure creating a wall stress in excess of the static yield stress is
applied to the gelled oil, initially only a narrow region of length Lf deforms and breaks down under stress. This yielded region is compressed by the entering ICF, which is also compressed. Eventually,
at time t = t0 the entire gelled oil plug will yield and move together
with the ICF at the same mass ow rate, as shown in Fig. 27.
For the calculations in this model from Davidson et al. (2004),
the bulk mass ow rate, G, is rst guessed (can use value from previous time step). Then the frictional factor, fk, is calculated by iteration for each longitudinal ICF and OGF segment at current time
step using the BuckinghamReiner equation for pipe ow of a
time-independent Bingham uid, and empirical relationships
developed by other authors for calculating the frictional factor in
laminar and turbulent ow. Equations (22) and (23) are used to
evaluate the mean velocity and shearing time, tsk,

qk Q k G
t sk t t 

22
k1
t0
M1

23

691

where k is the current subdivision out of M initial subdivisions of


the gelled oil used in the calculations; and qk is the dimensionless
density and Q k the dimensionless volumetric ow rate of the current subdivision. The pressure drop over each segment was calculated using following equation.

swk

fk qk t2k
Pk

4DLk
2

24

In the calculation procedure used by Davidson et al. (2004), the


length and density of each segment of oil is then updated to
account for the displacement of the OGF from the length of pipe
and the increase in the length of the ICF within the pipe. The ICF
rheology is assumed to be time-independent and it is therefore
separated into segments of equal length in each time step, with
the number of segments increasing by one with each time step.
The length of an ICF segment, DLICF, in time interval, i, at time
t  t 0 is given by following equation.

DLICF

P
OGF
LICF
DL  m
k1 Lk

K i
K i

25

Here K is the number of ICF segments chosen at time t = t0, m is


the number of remaining OGF segments within the pipe, and DLOGF
k
is the length of the k th gelled oil segment in the OGF for time
t  t 0 . Davidson et al. (2004) also calculated DLOGF
and the average
k
density for the k th segment in dimensionless form. Next in their
calculation procedure, the location of each segment and the ICF
OGF interface is determined relative to the downstream end of
the OGF plug. Then the pressure drop over each ICF and OGF
segment is summed to give the overall pressure drop, and the
difference between this value and the applied pressure drop

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

692

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

calculated and mass ow rate adjusted accordingly. This overall


process is iterated until the difference is negligible. G iterates to
zero in the case that the applied pressure is not high enough to
start ow at a given time, and the calculated pressure drop becomes the minimum required for start-up. The results of this model were signicantly different from those of the earlier model
developed by Chang et al. (1999), indicating the importance of fully
understanding the mechanism by which the gelled oil yields and is
displaced, and of determining the most realistic assumptions that
can be made during modelling.
Other researchers have also tackled understanding, modelling
and optimizing the restart of gelled lines. Borghi et al. (2003)
developed a model focusing on solid-like fracture propagation, viscous dissipation and compression of the broken gelled oil. Ekweribe et al. (2009), for instance, studied the effect of system pressure
on the restart of gelled subsea pipelines. They determined that
higher system pressures in subsea pipelines could lead to the formation of a weaker gel with lower yield strength, which would
mean that the necessary applied pressure for displacing it would
be more easily and cheaply achieved than might be predicted.

12. Conclusions
Contention still remains as to the specic mechanisms that govern wax deposition in pipelines. However, the importance of
molecular diffusion is generally accepted and shear dispersion is
usually not dismissed, at least due to the involvement of shear
forces in the removal of wax deposits, the accounting of which
has been shown by some authors to have a great impact on the
accuracy of wax deposition models. Many models have been developed based on the importance of these mechanisms, for which the
approach to a realistic representation of the solid phase wax components has a signicant impact on accuracy. Recently, a correct
heat-mass transfer analogy has been introduced into the modelling
of wax deposition, allowing for more accurate prediction across the
range of possible precipitation kinetics. In the future even more
accurate and robust models will be possible by combining this
new approach with an increased understanding of the mechanisms
involved in wax deposition and gelation and of the impact of other
species present in crude oil, such as asphaltenes and emulsied
water.
Understanding wax aging mechanisms is also very important to
fully understanding the process of the formation of wax deposits in
pipelines. Furthermore, understanding these mechanisms and predicting the CCN of particular crude oils would be helpful in determining what chemical inhibitors would be most effective for
preventing wax build-up in pipelines carrying those oils. The
continuing research into methods of inhibiting wax deposition
and removing deposits has the potential of making the maintenance of crude oil pipelines signicantly easier, as it becomes easier to optimize pigging frequency, to determine the minimum
pressure required to restart gelled lines, or even to avoid the need
for constant wax removal procedures by nding a way to costeffectively implement a promising method of control such as the
use of polar crude oil fractions or biological removal measures.

References
Addison, G.E., 1984. Parafn Control More Cost-Effective. In: SPE Eastern Regional
Meeting. Charleston, 1984. Society of Petroleum Engineers.
Al-Sabagh, A.M., El-Kafrawy, A.F., Khidr, T.T., El-Ghazawy, R.A., Mishrif, M.R., 2007.
Synthesis and evaluation of some novel polymeric surfactants based on
aromatic amines used as wax dispersant for waxy gas oil. J. Dispersion Sci.
Technol. 28, 976983.
Asperger, R.G., Sattler, R.E., Tolonen, W.J., Pitchford, A.C., 1981. Prediction of Wax
Buildup in 24 inch, Cold Deep Sea Oil Loading Line. USMS 10363.

Azarinezhad, R., Chapoy, A., Anderson, R., Tohidi, B., 2010. A wet cold-ow
technology for tackling offshore ow-assurance. SPE Projects, Facilities and
Construction 5, 5864.
Azevedo, L.F.A., Teixeira, A.M., 2003. A critical review of the modeling of wax
deposition mechanisms. Petrol. Sci. Technol. 21, 393408.
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
and Exhibition. Moscow, 2008. Society of Petroleum Engineers.
Banki, R., Hoteit, H., Firoozabadi, A., 2008. Mathematical formulation and numerical
modeling of wax deposition in pipelines from enthalpyporosity approach and
irreversible thermodynamics. Int. J. Heat Mass Transfer 51, 33873398.
Barnes, H.A., 1999. The Yield stressa review or pamsa qeieverything ows? J.
Non-Newtonian Fluid Mech. 81, 133178.
Bello, O.O., Fasesan, S.O., Teodoriu, C., Reinicke, K.M., 2006. An evaluation of the
performance of selected wax inhibitors on parafn deposition of Nigerian crude
oils. Pet. Sci. Technol. 24, 195206.
Bilderback, C.A., McDougall, L.A., 1963. Complete parafn control in petroleum
production. SPE J. Petrol. Technol. 21, 11511156.
Borghi, G.P., Correra, S., Merlini, M., Carniani, C., 2003. Prediction and scaleup of
waxy oil restart behavior. In: SPE International Symposium on Oileld
Chemistry. Houston, 2003. Society of Petroleum Engineers.
Bummer, B.L., 1971. Improved parafn prevention techniques reduce operating
costs, powder river basin, wyoming. In: Rocky Mountain Regional Meeting of
the Society of Petroleum Engineers of AIME. Billings, 1971. American Institute
of Mining, Metallurgical, and Petroleum Engineers.
Burger, E.D., Perkins, T.K., Striegler, J.H., 1981. Studies of wax deposition in the trans
alaska pipeline. J. Petrol. Technol. 33, 10751086.
Carmen Garca, M., Urbina, A., 2003. Effect of crude oil composition and blending on
owing properties. Petrol. Sci. Technol. 21, 863878.
Carmen Garca, M., 2001. Parafn deposition in oil production. In: SPE International
Symposium on Oileld Chemistry. Houston, 2001. Society of Petroleum
Engineers.
Carmen Garca, M., Orea, M., Carbognani, L., Urbina, A., 2001. The effect of parafnic
fractions on crude oil wax crystallization. Petrol. Sci. Technol. 19, 189196.
Chakrabarti, D.P., Das, G., Das, P.K., 2006. The transition from water continuous to
oil continuous ow pattern. AIChE J. 52, 36683678.
Chakrabarti, D.P., Das, G., Das, P.K., 2007. Identication of stratied liquidliquid
ow through horizontal pipes by a non-intrusive optical probe. Chem. Eng. Sci.
62, 18611876.
Chang, C., Boger, D.V., Nguyen, Q.D., 1998. The yielding of waxy crude oils. Ind. Eng.
Chem. Res. 37, 15511559.
Chang, C., Boger, D.V., Nguyen, Q.D., 2000. Inuence of thermal history on the waxy
structure of statically cooled waxy crude oil. SPE J. 5, 148157.
Chang, C., Nguyen, Q.D., Rnningsen, H.P., 1999. Isothermal start-up of pipeline
transporting waxy crude oil. J. Non-Newtonian Fluid Mech. 87, 127154.
Chen, X., Tsang, Y., Zhang, H.Q., Chen, T.X., 2007. Pressure-wave propagation
technique for blockage detection in subsea owlines. In: SPE Annual
Technical Conference and Exhibition. Anaheim, 2007. Society of Petroleum
Engineers.
Chen, X.T., Butler, T., Volk, M., Brill, J.P., 1997. Techniques for measuring wax
thickness during single and multiphase ow. In: SPE Annual Technical
Conference and Exhibition. San Antonio, 1997. Society of Petroleum
Engineers.
Cheng, D.C.H., Evans, F., 1965. Phenomenological characterization of the rheological
behavior of inelastic reversible thixotropic and antithixotropic uids. Brit. J.
Appl. Phys. 16, 15991617.
Cordoba, A.J., Schall, C.A., 2001. Application of a heat method to determine wax
deposition in a hydrocarbon binary mixture. Fuel 80, 12851291.
Correra, S., Fasano, A., Fusi, L., Merino-Garcia, D., 2007. Calculating deposit
formation in the pipelining of waxy crude oils. Meccanica 42, 149165.
Coutinho, J.A.P., Daridon, J.L., 2005. The limitations of the cloud point measurement
techniques and the inuence of the oil composition on its detection. Pet. Sci.
Technol. 23, 11131128.
Coutinho, J.A.P., 1998. Predictive UNIQAC: a new model for the description of
multiphase solid-liquid equilibria in complex hydrocarbon mixtures. Ind. Eng.
Chem. Res. 37, 4870.
Coutinho, J.A.P., Lopes-da-Silva, J.A.L., Ferreira, A., Soares, M.S., Daridon, J.L., 2003.
Evidence for the aging of wax deposits in crude oils by ostwald ripening. Petrol.
Sci. Technol. 21, 381391.
Davidson, M.R., Nguyen, Q.D., Chang, C., Rnningsen, H.P., 2004. A model for restart
of a pipeline with compressible gelled waxy crude oil. J. Non-Newtonian Fluid
Mech. 123, 269280.
de Oliveira, M.C.K., Carvalho, R.M., Carvalho, A.B., Couto, B.C., Faria, F.R.D., Cardoso,
R.L.P., 2010. Waxy crude oil emulsion gel: impact on ow assurance. Energy
Fuels 24, 22872293.
Duffy, D.M., Rodger, P.M., 2002. Modeling the activity of wax inhibitors: a case
study of poly(octadecyl acrylate). J. Phys. Chem. B 106, 1121011217.
Duffy, D.M., Moon, C., Rodger, P.M., 2004. Computer-assisted design of oil additives:
hydrate and wax inhibitors. Mol. Phys. 102, 203210.
Edmonds, B., Moorwood, T., Szczepanski, R., Zhang, X., 2008. Simulating wax
deposition in pipelines for ow assurance. Energy Fuels 22, 729741.
Ekweribe, C.K., 2008. Quiescent Gelation of Waxy Crudes and Restart of Shut-in
Subsea Pipelines (MS Thesis). Norman, Oklahoma: University of Oklahoma.
Ekweribe, C.K., Civan, F., Lee, H.S., Singh, P., 2009. Interim Report on Pressure Effect
on Waxy-Crude Pipeline-Restart Conditions Investigated by a Model System.
SPE Projects, Facilities & Construction, pp. 6174.

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694


Etoumi, A., El Musrati, I., El Gammoudi, B., El Behlil, M., 2008. The reduction of wax
precipitation in waxy crude oils by Pseudomonas species. J. Ind. Microbiol.
Biotechnol. 35, 12411245.
Farina, A., Fasano, A., 1997. Flow characteristics of waxy crude oils in laboratory
experimental loops. Math. Comput. Model. 25, 7586.
Fasano, A., Fusi, L., Correra, S., 2004. Mathematical models for waxy crude oils.
Meccanica 39, 441482.
Fielder, M., Johnson, R.W., 1986. The use of pour-point depressant additive in the
beatrice eld. In: SPE European Petroleum Conference. London, 1986. Society of
Petroleum Engineers.
Frigaard, I., Vinay, G., Wachs, A., 2007. Compressible displacement of waxy
crude oils in long pipeline startup ows. J. Non-Newtonian Fluid Mech.
147, 4564.
Fulford, R.S., 1975. Oilwell parafn prevention chemicals. In: SPE Regional Meeting.
Oklahoma, 1975. American Institute of Mining, Metallurgical, and Petroleum
Engineers.
Fung, G., Backhaus, W.P., McDaniel, S., Erdogmus, M., 2006. To pig or not to pig: the
marlin experience with stuck pig. In: Offshore Technology Conference. Houston,
2006. Offshore Technology Conference.
Fusi, L., Farina, A., 2004. A mathematical model for bingham-like uids with viscoelastic core. Z. Angew. Math. Phys. 55, 826847.
Fusi, L., 2003. On the stationary ow of a waxy crude oil with deposition
mechanisms. Nonlinear Anal. 53, 507526.
Groffe, D., Groffe, P., Takhar, S., Andersen, S.I., Stenby, E.H., Lindeloff, N. et al., 2001.
A wax inhibition solution to problematic elds: a chemical remediation process.
Petrol. Sci. Technol. 19, 205217.
Gudmundsson, J.S., Cold Flow Technology. In: Proc. 4th Intnl. Hydrates Conf.,
Yokohama (2002), pp. 912916.
Haghighi, H., Azarinezhad, R., Chapoy, A., Anderson, R., Tohidi, B., 2007. Hydraow:
avoiding gas hydrate problems. In: SPE Europe/EAGE Annual Conference and
Exhibition. London, 2007. Society of Petroleum Engineers.
Hammami, A., Ratulowski, J., Coutinho, J.A.P., 2003. Cloud points: can we measure
or model them? Petrol. Sci. Technol. 21, 345358.
Hammerschmidt, E.G., 1934. Formation of gas hydrates in natural gas
transportation lines. Ind. Eng. Chem. 26, 851855.
Haq, M.A., 1978. Deposition of Parafn Wax from its Solution with Hydrocarbons
(USMS 10541). Society of Petroleum Engineers.
He, Z., Mei, B., Wang, W., Sheng, J., Zhu, S., Wang, L. et al., 2003. A pilot test using
microbial parafn-removal technology in liaohe oileld. Petrol. Sci. Technol. 21,
201210.
Hernandez, O.C., Hensley, H., Sarica, C., Brill, J.P., Volk, M., Delle-Case, E., 2004.
Improvements in single-phase parafn deposition modeling. SPE Prod. Oper. 19,
237244.
Herri, J.-M., Pic, J.S., Gruy, F., Cournil, M., 1999. Methane hydrate crystallization
mechanism from in-situ particle size analyzer. AIChE J. 45, 590602.
Holder, G.A., Winkler, J., 1965. Wax crystallization from distillate fuels. J. Inst. Petrol.
51, 228252.
Houwink, R., 1958. Elasticity, Plasticity and Structure of Matter. Cambridge
University Press, New York.
Hunt, E.B.J., 1962. Laboratory study of parafn deposition. SPE J. Petrol. Technol. 14,
12591269.
Ilahi, M., 2005. Evaluation of Cold Flow Concepts (MS Thesis). Trondheim, Norway:
Norwegian University of Science and Technology.
Jang, Y.H., Blanco, M., Creek, J., Tang, Y., Goddard, W.A., 2007. Wax Inhibition
by comb-like polymers: support of the incorporation-perturbation
mechanism from molecular dynamics simulations. J. Phys. Chem. B 111,
1317313179.
Jennings, D.W., Breitigam, J., 2009. Parafn inhibitor formulations for different
application environments: from heated injection in the desert to extreme cold
arctic temperatures. Energy Fuels 24, 23372349.
Jorda, R.M., 1966. Parafn deposition and prevention in oil wells. SPE J. Petrol.
Technol. 18, 16051612.
Kelland, M.A., 2009. Production Chemicals for the Oil and Gas Industry. CRC
Press.
Kraynik, A.M., 1990. ER uid standards: comments on er uid rheology. In: Carlson,
J.D., Sprecher, A.F., Conrad, H. (Eds.), Proceedings of the Second International
Conference on ER Fluids. Raleigh, 1990. Technomic Publishing Company.
Kriz, P., Andersen, S.I., 2005. Effect of asphaltenes on crude oil wax crystallization.
Energy Fuels 19, 948953.
Lee, H.S., 2008. Computational and Rheological Study of Wax Deposition and
Gelation in Subsea Pipelines (PhD Dissertation). Ann Arbor, Michigan:
University of Michigan.
Lee, H.S., Singh, P., Thomason, W.H., Fogler, H.S., 2008. Waxy oil gel breaking
mechanisms: adhesive versus cohesive failure. Energy Fuels 22, 480487.
Leelavanichkul, P., Deo, M.D. Hanson, F.V., 2004. Crude oil characterization and
regular solution approach to thermodynamic modeling of solid precipitation at
low pressure. Petrol. Sci. Technol. 22, 973990.
Leiroz, A.T., Azevedo, L.F.A., 2005. Studies on the mechanisms of wax deposition in
pipelines. In: Offshore Technology Conference. Houston, 2005. Offshore
Technology Conference.
Li, H., Zhang, J., Yan, D., 2005. Correlations between the pour point or gel point and
the amount of precipitated wax for waxy crudes. Pet. Sci. Technol. 23, 1313
1322.
Lingelem, M.N., Majeed A.I., Stange, E. Industrial experience in evaluation of hydrate
formation, inhibition, and dissociation in pipeline design and operation. In:
Proc. 1st Intnl. Hydrates Conf., New-Paltz, NY (1993).

693

Lira-Galeana, C., Firoozabadi, A., Prausnitz, J.M., 1996. Thermodynamics of wax


precipitation in petroleum mixtures. AIChE J. 42, 239248.
Lockhart, T., Correra, S., 2005. Colloid and Surface Science Issues in the Petroleum
Industry. In: Ente Nazionale Idrocarburiz & Istituto della Enciclopedia Italiana
Encyclopaedia of Hydrocarbons. Roma, Italy: Istituto della Enciclopedia Italiana.
Ch. 3.2.1. pp. 167177.
Lopes-da-Silva, J.A., Coutinho, J.A.P., 2007. Analysis of the isothermal structure
development in waxy crude oils under quiescent conditions. Energy Fuels 21,
36123617.
Majeed, A., Bringedal, B., Overa, S., 1990. Model calculates wax deposition for n. sea
oils. Oil Gas J. 88, 6369.
Manka, J.S., Ziegler, K.L., 2001. Factors affecting the performance of crude oil waxcontrol additives. In: SPE Production and Operations Symposium. Oklahoma,
2001. Society of Petroleum Engineers.
Manka, J.S., Magyar, J.S., Smith, R.P., 1999. A novel method to winterize traditional
pour point depressants. In: SPE Annual Technical Conference and Exhibition.
Houston, 1999. Society of Petroleum Engineers.
Matzain, A., 1999. Multiphase Flow Parafn Deposition Modeling (PhD
Dissertation). Tulsa, Oklahoma: The University of Tulsa.
Matzain, A., Apte, M.S., Zhang, H., Volk, M., Brill, J.P., Creek, J.L., 2002. Investigation
of parafn deposition during multiphase ow in pipelines and wellborespart
1: experiments. J. Energy Res. Technol. 124, 180186.
Mendell, J.L., Jessen, F.W., 1970. Mechanism of Inhibition of Parafn Deposition in
Crude Oil Systems. In: Ninth Bienniel Production Techniques Symposium.
Wichita Falls, 1970. American Institute of Mining, Metallurgical and Petroleum
Engineers.
Merino-Garcia, D., Correra, S., 2008. Cold ow: a review of a technology to avoid
wax deposition. Pet. Sci. Technol. 26, 446459.
Merino-Garcia, D., Margarone, M., Correra, S., 2007. Kinetics of waxy gel formation
from batch experiments. Energy Fuels 21, 12871295.
Mersmann, A. 2002. Crystallization Technology Handbook, second ed. Marcel
Dekker.
Nautiyal, S.P., Kumar, S., Srivastava, S.P., 2008. Crystal structure of n-parafn
concentrates of crude oils. Pet. Sci. Technol. 26, 13391346.
Newberry, M.E., Barker, K.M., 1985. Formation damage prevention through the
control of parafn and asphaltene deposition. In: SPE Production Operations
Symposium. Oklahoma City, 1985. Society of Petroleum Engineers.
Oh, K., Jemmett, M., Deo, M., 2009. Yield behavior of gelled waxy oil: effect of stress
application in creep ranges. Ind. Eng. Chem. Res. 48, 89508953.
Paso, K.G., Fogler, H.S., 2003. Inuence of n-parafn composition on the aging of
wax-oil gel deposits. AIChE J. 49, 32413252.
Paso, K.G., 2005. Parafn Gelation Kinetics (PhD Dissertation). Ann Arbor, Michigan:
University of Michigan.
Paso, K.G., Kompalla, T., Aske, N., Rnningsen, H.P., ye, G., Sjblom, J., 2009a. Novel
surfaces with applicability for preventing wax deposition: a review. J.
Dispersion Sci. Technol. 30, 757781.
Paso, K.G., Kompalla, T., Oschmann, H.J., Sjblom, J., 2009b. Rheological degradation
of model waxoil gels. J. Dispersion Sci. Technol. 30, 472480.
Paso, K.G., Silset, A., Srland, G., Gonalves, M.A.L., Sjblom, J., 2009c.
Characterization of the formation, ow ability, and resolution of Brazilian
crude oil emulsions. Energy Fuels 23, 471480.
Patton, C.C., 1970. parafn deposition from rened wax-solvent systems. Soc.
Petrol. Eng. J., 1724.
Pedersen, K.S., Rnningsen, H.P., 2003. Inuence of wax inhibitors on wax
appearance temperature, pour point, and viscosity of waxy crude oils. Energy
Fuels 17, 321328.
Petrellis, N.C., Flumerfelt, R.W., 1973. Rheological behavior of shear degradable oils:
kinetic and equilibrium properties. Can. J. Chem. Eng. 51, 291301.
Prausnitz, J.M., Lichtenthaler, R.N., Azevedo, E.G., 1986. Molecular Thermodynamics
of Fluid-phase Equilibria. Prentice-Hall, Englewood Cliffs, NJ.
Quenelle, A., Gunaltun, M., 1987. Comparison between thermal insulation coatings
for underwater pipelines. In: Offshore Technology Conference. Houston, 1987.
Offshore Technology Conference.
Raj, T.S., Chakrabarti, D.P., Das, G., 2005. Liquid-liquid stratied ow through
horizontal conduits. Chem. Eng. Technol. 28, 899907.
Ramrez-Jaramillo, E., Lira-Galeana, C., Manero, O., 2001. Numerical Simulation of
Wax Deposition in Oil Pipeline Systems. Petrol. Sci. Technol. 19, 143156.
Ramrez-Jaramillo, E., Lira-Galeana, C., Manero, O., 2004. Modeling wax deposition
in pipelines. Petrol. Sci. Technol. 22, 821861.
Rana, D.P., Bateja, S., Biswas, S.K., Kumar, A., Misra, T.R., Lal, B., 2010. Novel
microbial process for mitigating wax deposition in down hole tubular and
surface ow lines. In: SPE Oil and Gas India Conference. Mumbai, 2010. Society
of Petroleum Engineers.
Rao, B.M.A., Mahajan, S.P., Khilar, K.C., 1985. A Model on the breakdown of crude oil
gel. Can. J. Chem. Eng. 63, 170172.
Reistle, C.E.J., 1928. Methods of Dealing with Parafn Troubles Encountered in
Producing Crude Oils. USBM Technical Papers.
Reistle, C.E.J., 1932. Parafn and Congealing Oil Problems. USBM Bulletins.
Sarmento, R.C., Ribbe, G.A.S., Azevedo, L.F.A., 2004. Wax blockage removal by
inductive heating of subsea pipelines. Heat Transfer Eng. 25, 212.
Senra, M., Panacharoensawad, E., Kraiwattanawong, K., Singh, P., Fogler, H.S., 2008.
Role of n-alkane polydispersity on the crystallization of n-alkanes from
solution. Energy Fuels 22, 545555.
Senra, M., Scholand, T., Maxey, C., Fogler, H.S., 2009. Role of polydispersity and
cocrystallization on the gelation of long-chained n-alkanes in solution. Energy
Fuels 23, 59475957.

694

A. Aiyejina et al. / International Journal of Multiphase Flow 37 (2011) 671694

Sifour, M., Al-Jilawi, M.H., Aziz, G.M., 2007. Emulsication properties of


biosurfactant produced from Pseudomonas aeruginosa RB 28. Pak. J. Biol. Sci.
10, 13311335.
Singh, P., Venkatesan, R., Fogler, H.S., 2000. Formation and aging of incipient thin
lm waxoil gels. AIChE J. 46, 10591074.
Singh, P., Venkatesan, R., Fogler, H.S., 2001a. Morphological evolution of thick wax
deposits during aging. AIChE J. 47, 618.
Singh, P., Youyen, A., Fogler, H.S., 2001b. Existence of a critical carbon number in the
aging of a wax-oil gel. AIChE J. 47, 21112124.
Singhal, H.K., Sahai, G.C., Pundeer, G.S., Chandra, K., 1991. Designing and selecting
wax crystal modier for optimum eld performance based on crude oil
composition. In: Annual Technical Conference and Exhibition of the Society of
Petroleum Engineers. Dallas, 1991. Society of Petroleum Engineers.
Smith, P.B., Ramsden, R.M.J., 1978. The prediction of oil gelation in submarine
pipelines and the pressure required for restarting ow. In: European Offshore
Petroleum Conference and Exhibition. London, 1978. European Offshore
Petroleum Conference and Exhibition.
Solaimany Nazar, A.R., Dabir, B., Islam, M.R., 2005a. A multi-solid phase
thermodynamic model for predicting wax precipitation in petroleum
mixtures. Energy Sources 27, 173184.
Solaimany Nazar, A.R., Dabir, B., Islam, M.R., 2005b. Experimental and mathematical
modeling of wax deposition and propagation in pipes transporting crude oil.
Energy Sources 27, 185207.
Soni, H.P., Kiranbala, Bharambe, D.P., 2008. Performance-based design of wax
crystal growth inhibitors. Energy Fuels 22, 39303938.
Tinsley, J.F., Prudhomme, R.K., Guo, X., Adamson, D.H., Shao, S., Amin, D. et al., 2007.
Effects of polymers on the structure and deposition behavior of waxy oils. In:
SPE International Symposium on Oileld Chemistry. Houston, 2007. Society of
Petroleum Engineers.
Torres, G.A., Turner, C., 2005. Method of straight lines for a bingham problem as a
model for the ow of waxy crude oils. Electron. J. Diff. Equat., 115.

Venkatesan, R., Fogler, H.S., 2004. Comments on analogies for correlated heat and
mass transfer in turbulent ow. AIChE J. 50, 16231626.
Venkatesan, R., stlund, J., Chawla, H., Wattana, P., Nydn, M., Fogler, H.S., 2003. The
Effect of asphaltenes on the gelation of waxy oils. Energy Fuels 17, 16301640.
Vinay, G., Wachs, A., Frigaard, I., 2007. Start-up transients and efcient computation
of isothermal waxy crude oil ows. J. Non-Newtonian Fluid Mech. 143, 141
156.
Visintin, R.F.G., Lockhart, T.P., Lapasin, R., DAntona, P., 2008. Structure of waxy
crude oil emulsion gels. J. Non-Newtonian Fluid Mech. 149, 3439.
Wang, K.S., Wu, C.H., Creek, J.L., Shuler, P.J., Tang, Y., 2003. Evaluation of effects of
selected wax inhibitors on parafn deposition. Petrol. Sci. Technol. 21, 369379.
Wang, Q., Sarica, C. & Volk, M., 2008. An experimental study on wax removal in
pipes with oil ow. J. Energy Resour. Technol., 130.
Wardhaugh, L.T., Boger, D.V., 1991. The measurement and description of the
yielding behavior of waxy crude oil. J. Rheol. 35, 11211156.
Wardhaugh, L.T., Boger, D.V., Tonner, S.P., 1988. Rheology of waxy crude oils. In: SPE
International Meeting on Petroleum Engineering. Tianjin, 1988. Society of
Petroleum Engineers.
Woo, G.T., Garbis, S.J. & Gray, T.C., 1984. Long-term control of parafn deposition.
In: Technical Conference and Exhibition. Houston, 1984. Society of Petroleum
Engineers.
Wuhua, C., Zongchang, Z., 2006. Thermodynamic modeling of wax precipitation in
crude oils. Chin. J. Chem. Eng. 14, 685689.
Yang, X., Kilpatrick, P., 2005. Asphaltenes and waxes do not interact synergistically
and coprecipitate in solid organic deposits. Energy Fuels 19, 13601375.
Zaman, M., Agha, K.R., Islam, M.R., 2006. Laser based detection of parafn in crude
oil samples: numerical and experimental study. Pet. Sci. Technol. 24, 722.
Zaman, M., Bjorndalen, N., Islam, M.R., 2004. Detection of precipitation in pipelines.
Petrol. Sci. Technol. 22, 11191141.

You might also like