Sie sind auf Seite 1von 9

Chemical Engineering Science 97 (2013) 311–319

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Science


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

A model of wax deposit layer formation


Dmitry Eskin n, John Ratulowski, Kamran Akbarzadeh
Schlumberger DBR Technology Center, 9450-17 Avenue, Edmonton, AB, Canada T6N 1M9

H I G H L I G H T S

 We model wax deposit layer formation.


 We assume thermodynamic equilibrium between wax solids and a fluid.
 We consider a growing porous layer with a variable porosity.
 Two major model equations describe wax diffusion and deposit growth rate.
 We numerically study model performance varying different parameters.

art ic l e i nf o a b s t r a c t

Article history: A model of wax deposit layer formation is developed. The model is based on a mathematical description
Received 31 January 2013 of diffusion of wax molecules through a growing porous layer. The deposit layer porosity reduction is
Received in revised form caused by the growth of wax crystals, composing porous medium, in a wax-saturated hydrocarbon fluid.
15 April 2013
The two major model equations are: (1) diffusion of wax molecules through a porous medium with
Accepted 18 April 2013
varying porosity; and (2) deposit layer growth rate. The model takes into account a diffusivity reduction
Available online 27 April 2013
with a porosity decrease and a shear-removal effect. The model performance is illustrated by
Keywords: computations of the wax deposition in a laboratory Taylor–Couette device.
Deposition & 2013 Elsevier Ltd. All rights reserved.
Diffusion
Modeling
Porous media
Thermal conductivity
Wax

1. Introduction The pigging procedure is applied periodically and the frequency


depends on a deposit growth rate forecast. To properly schedule
Wax deposition usually occurs in pipelines when the tempera- the pigging operation, an accurate modeling of the deposition
ture in the near-wall area drops below the so-called wax appear- phenomenon is required.
ance temperature (WAT). Precipitation of wax particles from a It is known that the solid wax content in the deposit layer
fluid (the lower the temperature, the higher the particle concen- increases with time. This change of a wax deposit layer structure is
tration) causes a corresponding deficiency of wax molecules in the usually referred to as aging. Nevertheless, many conventional
fluid. The wax compositional gradient leads to a diffusive flux models (e.g., Svendsen, 1993; Akbarzadeh and Zougari, 2008)
toward the wall, which causes the deposit layer formation. Thus, assume that a deposit layer grows maintaining mean solids
the wax deposition is a temperature gradient driven process: content during the entire production time. Although, the wax
i.e., zero gradient guarantees zero concentration gradient that deposit analysis shows that this mean value may vary in a wide
means no deposition. range (0.05–0.4 according to Akbarzadeh and Zougari (2008) and
The deposit layer, growing with time, can cause a significant Hernandez et al. (2003)), the constant mean solids content
reduction in a pipe cross-section and a flow rate respectively. assumption allows obtaining a reasonably good agreement
Eventually, the deposit can almost entirely obstruct the pipe cross- between simulated and experimental deposition rates under
section; therefore, under field conditions, the deposit layer formed laboratory conditions. Note that a duration of a regular wax
on the pipe wall is usually removed mechanically (by pigging). deposition experiment is, usually, not sufficient to reveal a
significant effect of the aging phenomenon on the deposition rate.
Thus, it is important to emphasize works of Singh et al. (2000,
n
Corresponding author. Tel.: +1 780 577 1311; fax: +1 780 450 1668. 2001), who modeled the aging process based on a description
E-mail address: deskin@slb.com (D. Eskin). of wax molecular diffusion through the porous deposit layer.

0009-2509/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ces.2013.04.040
312 D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319

A comprehensive aging model of Singh et al. (2001) assumed that fluid to the deposit layer surface:
the diffusion mass flux from a bulk fluid to the outer deposit layer
∂c
surface is artificially separated to two components. The first is qd ¼ −D ð1Þ
∂n
spent on an increase in the deposit thickness, while the second,
the diffusion flux of wax of molecules inside the layer, causes an where c is mass concentration of wax molecules in a hydrocarbon
increase in the solids fraction due to the adsorption of wax fluid, D is the diffusivity of wax molecules in a fluid, ∂c/∂n is the
molecules from a fluid. While the model concept looks plausible, concentration gradient at the deposit layer surface, n is the
the practical realization was poor. The authors assumed zero coordinate axis, normal to the layer surface.
molecular diffusion flux at the wall and used it as the boundary Based on experimental data for a mass fraction of precipitated
condition. From a physical standpoint, this condition contradicts to wax crystals vs. temperature, Burger et al. (1981) suggested
a basic assumption employed by Singh et al. (2001), which states calculating the wax concentration gradient ∂c/∂n assuming the
that the wax deposition is a thermally driven process and the thermodynamic equilibrium between precipitating wax solids and
diffusive flux is proportional to the heat flux. Because the wax a hydrocarbon fluid. Then, the local wax concentration is unam-
deposition is a thermally driven process, the heat flux within the biguously determined by the local temperature and, therefore, the
deposit layer and at the wall cannot be zero; therefore, the concentration gradient is proportional to the temperature gradient
diffusion flux cannot be zero as well. The authors did not also and calculated as follows (Burger et al., 1981):
take into account an important shear-removal phenomenon, ∂c dc ∂T
caused by a partial removal of the deposit material due to the ¼ ð2Þ
∂n dT ∂n
fluid-deposit surface viscous friction. The model deficiencies and
especially employment contradicting boundary condition pre- where dc/dT is the rate of a wax concentration increase with
vented Singh et al. (2001) from obtaining valid results. These temperature (the mass fraction of precipitated wax crystals
authors computed the deposit layer growth and obtained an decreases with the same rate) and ∂T/∂n is the temperature
unexpected layer structure: the solid wax fraction decreased from gradient.
the deposit layer surface toward the wall. Singh et al. (2001) could The thermodynamic equilibrium assumption is currently
not clearly explain the mechanism of deposit aging leading to the employed by majority of wax deposition modelers (e.g.,
layer structure obtained. Svendsen, 1993; Akbarzadeh and Zougari, 2008). However, it is
For practical purposes Hernandez et al. (2003) and Huang et al. important to mention a work of Lee (2007), who developed a
(2011) recommend using a simplified approach of Singh et al. model of wax deposition accounting for precipitation kinetics of
(2000), who formulated the equations describing an evolution of wax molecules in the boundary layer. That model accounts for an
the solids fraction averaged over the deposit layer thickness with effect of fluid supersaturation with wax and gives a higher wax
time. However, such an approach cannot be justified because the deposition rate than the model based on the solids–liquid equili-
rate of a solids content increase within the deposit layer strongly brium assumption. In the current paper we assume the thermo-
depends on the diffusivity of wax molecules through the porous dynamic equilibrium between the phases.
layer that in turn is a strong function of the solids fraction itself. The deposit layer formation process to a great extent deter-
Thus, the major target of the current work is development of a mines the major deposition characteristics, such as the deposition
physically justified model of the deposit aging. rate and the deposit thickness. A simplified deposit layer diagram
We would like to also emphasize that matching wax deposition is shown in Fig. 1. Let us assume that the deposit layer has a
simulation results to experimental data either in a Couette device porous structure with a varying porosity. We will limit our study to
or a flow loop is, in general, a simple task. This observation is the cases, in which the deposit thickness is significantly smaller
explained by a sufficient (often excessive) number of fitting than the pipe radius. Following the majority of the previous
parameters; therefore, practically all published models (e.g., models, we also assume that the layer is immobile; i.e., solid,
Burger et al., 1981; Hernandez et al., 2003; Akbarzadeh and while its growth is decelerated by a partial removal of the
Zougari, 2008) demonstrated very good performance for given deposited material due to a viscous friction of a fluid flow with
experimental data. However, only a proper description of a the deposit surface. Let us formulate an equation of wax balance
deposition mechanism guarantees proper deposition forecasting for the deposit layer, composed of solid wax particles and waxy
on a field scale. The modeling approach that we present in this fluid, as a diffusion equation for a porous medium:
work is based on the same major verified principles as its ∂½cφ þ ρwax ð1−φÞ ∂ ∂c
predecessors (e.g., Svendsen, 1993; Akbarzadeh and Zougari, ¼ Dm ð3Þ
∂t ∂x ∂x
2008). However, a robust and self-consisted description of the
where Dm is the diffusivity of wax molecules through the porous
aging phenomenon is expected to improve wax deposition fore-
medium, x is the coordinate axis normal to the solid surface (x ¼0
casting under field conditions.
at the surface), φ is the porosity, ρwax is the density of the
solid wax.
The diffusivity Dm is a function of the wax molecular diffusivity
2. Model
in a fluid and the porosity. Singh et al. (2000, 2001) and other
investigators (Hernandez et al., 2003), who modeled the wax
2.1. Discussion of model assumptions and derivation of equations
aging, calculated the diffusivity through porous media by a semi-
empirical equation of Cussler et al. (1988) derived for porous
Currently, a majority of researchers assume the molecular
diffusion is the major wax deposition mechanism. This approach
was suggested by Burger et al. (1981), who included in their model qd qsr
also the deposition of precipitated wax particles on the wall.
However, in later publications, majority of authors successfully x, z
modeled the wax deposition employing molecular diffusion as the
only deposition mechanism (e.g., Svendsen, 1993; Huang et al., δ
2011). Thus, the total wax deposition rate can be accurately
calculated as the diffusive mass flux of wax molecules from a bulk Fig. 1. Deposit layer diagram.
D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319 313

media composed of flake-like particles. Cussler's equation is a the deposit thermal conductivity on porosity into the model, an
function of the wax particle aspect ratio, which Singh et al. (2000, increase in the layer conductivity during the deposit aging can be
2001) and Hernandez et al. (2003) used as a deposition model approximately accounted as follows. The thermal conductivity of
fitting parameter. Based on fitting the deposition modeling results the deposit, as a thermal insulation layer, is assumed to be
to the experimental data, Singh et al. (2000) claimed that the constant but depends on the mean layer porosity and any reason-
aspect ratio parameter significantly reduces with a decrease in the able correlation to calculate λ(φm) can be employed.
fluid flow rate. In our opinion, this aspect ratio behavior could also The condition d lnλ=dφ ¼ 0 yields ∂2c/∂x2 ¼0 that leads to the
indicate a poor applicability of the Cussler's equation to descrip- constant concentration gradient∂c/∂x across the deposit layer.
tion of the diffusivity through the wax layer. To calculate Dm in this Thus, Eq. (5) takes the following form:
work we will employ a well-known Archie law that is suitable for ∂φ ∂φ ∂c
modeling of the diffusivity through majority of porous materials ðρwax −cm Þ ¼ −nDφn−1 ð6Þ
∂t ∂x ∂x
(Peng et al., 2012):
where cm is the mean equilibrium wax concentration in the fluid
Dm ¼ Dφn ð4Þ that fills the deposit pores.
where n is the empirical parameter (n≥1). Because the thermal diffusivities of the fluid and the deposit
After substituting Eq. (4) into Eq. (3) we obtain material are similar, the temperature gradient at the deposit layer
surface on the fluid side is nearly equal to that on the deposit side;
∂φðc−ρwax Þ ∂φ ∂c ∂2 c therefore, the corresponding wax concentration gradients are also
¼ nDφn−1 þ Dφn 2 ð5Þ
∂t ∂x ∂x ∂x equal. Then, the concentration gradient, which is constant over the
The wax concentration gradient is calculated by Eq. (2). layer thickness, can be expressed through the molecular flux qd to the
According to the experimental data (e.g., Akbarzadeh and top surface of the deposit layer (see Eq. (1)) as dc/dx¼−qd/D. Substi-
Zougari, 2008), the function dc/dT slowly changes with the tuting this gradient into Eq. (6) we obtain the following equation:
temperature and its absolute value is relatively small (dc/ ∂φ 1 ∂φ
¼− nφn−1 qd ð7Þ
dT∼10−2–10−1 kg/m3 K). Therefore, in practical applications, both ∂t ρwax −cm ∂x
the function dc/dT and the equilibrium wax concentration c can be
Note that the expected “plus” in front of the right-hand side of
assumed constant over the layer thickness and during the entire
Eq. (7) was omitted because although the deposit flux qd is
time of the deposit layer formation. It is also important to
negative in the employed coordinate system, we will use its
emphasize that the thermal diffusivity of the deposit layer a¼ λ/
absolute value in our further analysis.
cpρd∼10−7 m2/s (λ is the layer thermal conductivity, cp is the heat
Because the layer thickness increases with time, it is conve-
capacity per unit mass and ρd is the deposit density) is much larger
nient to analyze the deposit growth in a moving coordinate
than the molecular diffusivity of wax (D∼10−9 m2/s). Therefore, the
system. Instead of x we will employ the dimensionless coordinate
heat conduction is much faster (by factor of a/D∼100) process than
z¼ x/δ, where δ is the time-dependent deposit layer thickness. Any
the molecular diffusion. Because the wax mass transport across
derivative in respect to time in a moving coordinate system is
the deposit layer, determining the porosity profile evolution, is
 
entirely controlled by the molecular diffusion, the temperature ∂ ∂ dδ=dt ∂
¼ þz ð8Þ
and the corresponding equilibrium wax concentration profiles can ∂t mov ∂t δ ∂z
be considered steady-state (∂T/∂t ¼∂c/∂t¼0) for Eq. (5). From the
Eq. (7) is written as
steady-state equation of heat conduction (e.g., Holman, 1990),    
formulated for the deposit layer, one can easily express the second ∂φ ∂φ 1 qd dδ
¼− nφn−1 −z ð9Þ
derivative of the temperature in respect of the coordinate as ∂t mov ∂z δ ρwax −cm dt
∂2 T=∂x2 ¼ −ðd lnλ=dφÞð∂φ=∂xÞð∂T=∂xÞ. Then, using Eq. (2) the second
Let us calculate the deposit layer growth rate from a wax mass
derivative of the wax concentration can be written as
balance equation, formulated for the layer of the thickness δ, as
∂2 c=∂x2 ¼ −ðd lnλ=dφÞð∂φ=∂xÞð∂c=∂xÞ. Literature search showed that
follows:
there are no published data about dependence of the wax deposit
Z
thermal conductivity on the porosity; i.e. calculating the term ∂ δ
qd −qsr ¼ ðcφ þ ρwax ð1−φÞÞdx ð10Þ
d lnλ=dφ is problematic. The thermal conductivities of oil (λf) and ∂t 0
solid wax (λwax) vary in wide ranges. According to Ukrainczyk et al.
where qsr ¼εδ is the shear-removal rate and ε is the proportionality
(2010), differences between thermal conductivities of solid waxes,
coefficient.
reported in different studies, exceed 100% (λwax∼0.15–0.35 W/m K).
Note that the shear-removal term qsr expresses a simple idea of
Oil conductivities are usually lower (e.g., Elam et al., 1989), being
Kern and Seaton (1959), who assumed that the removal rate is propor-
in the range of λf∼0.1–0.2 W/m K. In practice, the conductivity of
tional to the layer thickness. This approach is regularly employed
solid wax does not significantly differ from the conductivity of
for wax deposition modeling (e.g., Akbarzadeh and Zougari, 2008).
hydrocarbon fluid (frequently, λwax∼0.2 and λf∼0.14 W/m K). From
The employed shear-removal term is obviously oversimplified. It does
practical standpoint, it makes sense to assume the constant
not even contain the porosity that determines the porous layer
conductivity of the deposit (d lnλ=dφ ¼ 0). There are the two
strength and the removed mass. Unfortunately, there are no systema-
reasons for this assumption. The first is based on the fact,
tic data on the shear-removal in the literature. Some investigators (e.g.,
mentioned in Introduction: the mean deposit layer porosity during
Hernandez et al., 2003) claim that qsr can be even assumed constant
the deposit growth usually varies not in the very wide range
over the entire deposition process. Therefore, we employed the shear-
(φm≈0.6–0.95). The second reason is associated with a very poor
removal term that has been most frequently used by modelers.
accuracy of correlations for calculating thermal conductivity of
After transforming Eq. (10) into the moving coordinate system,
porous materials. The low accuracy is caused by a strong depen-
we obtain:
dence of the conductivity on a pore structure (Carson et al., 2005),
while experimental data for the wax deposit conductivity are not R1   R1  
qd −εδ ¼ ∂
0 ∂t mov ðcφ þ ρwax ð1−φÞÞδ dz ¼ − 0 ∂φ
∂t mov ðρwax −cm Þδdz
available; therefore, an applicability of any known correlation to Z 1
the wax deposit layer cannot be well justified. To avoid unneces- dδ
þ ðcm φ þ ρwax ð1−φÞÞdz ð11Þ
sary complexities, associated with introducing the dependence of dt 0
314 D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319

Note that furthermore we will assume the moving coordinate Modeling the deposition in an entire pipe requires meshing the
system and omit the subscript “mov”. computational domain over the pipe axis ξ and solving Eqs. (14)
As a result of substituting Eq. (9) into Eq. (11) and performing and (15) for each computational cell. A transport pipeline can be
integration over the normalized deposit layer thickness, we obtain long (up to tens of kilometers), i.e., a large number of computa-
the equation for the deposit layer growth rate as follows: tional cells can be required; therefore, practical application of the
model developed is associated with relatively large computational
dδ ðq =ðρ −cm ÞÞð1−φð1Þn þ φð0Þn −κΘÞ
¼ d wax ð12Þ arrays.
dt ðρwax =ðρwax −cm ÞÞ−φð1Þ
where Θ ¼[(ρwax−cm)/qd]δ is the characteristic time and κ ¼[ε/
(ρwax−cm)] is the shear-removal constant.
Note that to link the shear-removal term with the shear stress 3. Identification of model parameters
at the wall the coefficient κ is often presented as follows (Kern and
Seaton, 1959): 3.1. Model parameters and their ranges
κ ¼ ksr τw ð13Þ
For identification of model parameters it makes sense employ-
where ksr is the shear-removal factor that we will use as the model
ing a relatively short flow loop, in which variations of the fluid
parameter instead of κ, τw is the shear stress at the wall.
temperature and the corresponding deposit thickness are small
Taking into account Eq. (12), Eq. (9) can be written as follows:
  along the pipe axis. A majority of researchers use a flow loop for
∂φ ∂φ 1 d lnδ wax deposition experiments. For modeling an evolution of the
¼− nφn−1 −zΘ ð14Þ
∂t ∂z Θ dt deposit layer in a device, where the deposit layer thickness
changes only with time, we will rewrite Eqs. (14) and (15) in a
where
dimensionless form as
d lnδ 1 1−φð1Þn þ φð0Þn −κΘ
¼ ð15Þ  
dt Θ υ−φð1Þ ∂φ ∂φ d lnδ
¼− nφn−1 −z ð21Þ
∂τ ∂z dτ
and υ ¼ ρwax =ðρwax −cm Þ:
The set of Eqs. (14) and (15) describes the problem with the where
initial conditions. The initial condition for Eq. (14) is the porosity
distribution at the initial time moment: d lnδ 1−φð1Þn þ φð0Þn −κΘ
¼ ð22Þ
dτ υ−φð1Þ
φð0; zÞ ¼ f ðzÞ ð16Þ
The initial condition for Eq. (15) is the initial layer thickness: τ ¼t/Θ is the dimensionless time.
Eq. (21), coupled with Eq. (22), can be straightforwardly
δð0Þ ¼ δ0 ð17Þ integrated by a finite difference method.
The trivial solution of the problem is obtained for f(z)¼const. The initial porosity distribution over the deposit layer thick-
In this case, the porosity does not change with time. ness, needed as the initial condition to solve Eq. (21), is deter-
mined by a complicated physics and unknown. It follows from
2.2. Application of the model to a pipe flow Eq. (21) that because the mean layer porosity decreases with time,
the initial porosity has to increase from the bottom to the upper
We would like to note that an application of the model boundary. Further, as an engineering estimation we will assume
developed to a pipe flow is straightforward but can be computa- the linear initial porosity distribution not significantly deviating
tionally expensive. The deposit layer growth at a certain location from a uniform one. The initial boundary layer thickness δ(0) that
along a pipe is calculated by Eqs. (14) and (15). The only variable is the boundary condition for Eq. (22) must be chosen small to
that needs to be calculated by a separate algorithm is the minimize its effect on the solution. In practice it has, probably, to
molecular flux of wax molecules to the deposit layer that is be comparable to the pipe wall roughness.
calculated based on Eqs. (1) and (2) as follows: A simple analysis of Eq. (21) shows that if the exponent n
equals 1 or 2, the initially linear porosity distribution changes with
dc jqj
qd ¼ D ð18Þ time but remains linear. In these cases, the derivative of the
dT λf porosity with respect to the dimensionless thickness is calculated
where q ¼−λf(∂T/∂x)surf is the heat flux to the deposit surface, and λf as ∂φ/∂z¼ φ(τ,1)−φ(τ,0) and the porosity evolution at the lower and
is the fluid thermal conductivity. the upper surfaces of the layer is described by the ordinary
The heat flux is determined as (e.g., Holman, 1990): differential equations that follow from Eq. (21).
To use the deposition model developed, we need to set five
q ¼ kbo ðT b −T o Þ ð19Þ
major parameters: the porosities φ(0,0), φ(0,1), the exponent n, the
where kbo is the heat transfer coefficient from a fluid inside the molecular diffusion constant kmd needed to properly define the
pipe to an environment, Tb is the bulk (mean) fluid temperature diffusivity of wax molecules D (see Eq. (23) below) and the shear-
inside of a pipe, To is the ambient temperature outside of a pipe. removal coefficient κ (or ksr in Eq. (13)). All of these parameters
The bulk temperature distribution along a pipe can be serve as fitting coefficients. Due to absence of any data about the
described by the simple ordinary differential equation derived initial deposit layer structure, we will assume that the initial
from a heat balance (e.g., Eskin et al., 2009): porosity φ(0,1) is determined as a maximum porosity that allows
maintaining the layer integrity under action of the shear stress.
dT b 4k ðT −T o Þ
¼ − bo b ð20Þ Based on the experimental data on the mean layer porosity,
dξ ρf cp UDp
mentioned in Section 1, we can conclude that the porosity φ(0,1)
where cp is the fluid heat capacity, Dp is the inner pipe diameter, has to be lower than 0.95. Let us assume the possible range of the
U is the mean flow velocity, ξ is the axial coordinate. exponent n to be from 1 to 5 that covers the majority of porous
The heat transfer coefficient kbo can be determined by using a materials (Peng et al., 2012). The diffusivity of wax molecules D
standard technique (Holman, 1990). can be calculated by the empirical correlation as follows (Hayduk
D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319 315

and Minhas, 1982): after that, the expression in the parenthesis on the right-hand side
of the resulting equation is equated to zero at z ¼1. Then, the
D ¼ kmd T 1:47 μ10:2=υw −0:791 υ−0:71 ð23Þ
w coefficient κ is obtained in the following form:
where kmd is the molecular diffusion constant that needs to be
tuned experimentally, μ is the fluid viscosity in centipoises, and υw 1−φðτ; 1Þn þ φðτ; 0Þn −nφðτ; 1Þn−1 ðυ−φðτ; 1ÞÞ
κ¼ ð24Þ
is the wax molar volume in cubic centimeters per mole. Θ
A procedure of identification of the shear-removal coefficient κ At a certain layer thickness, the coefficient κ becomes equal to
is discussed in the next section. κt, corresponding to the 2nd stage of the shear-removal process,
We would like to emphasize that to evaluate the model and further the coefficient κt remains constant. The coefficient κt
parameters, both the deposit mass and the mean solids content can be identified by fitting a computed layer thickness to a
in the deposit have to be accurately determined for each experi- measured one for an extended period of time.
mental run. Experiments required for reliable identification of the Note that due to introduction of the shear-removal stages into
deposit aging model parameters have to be noticeably longer (tens the model, the distributions of the modeled layer parameters over
of hours) than those (usually, several hours long) needed to time can be characterized by visible cusps at the switching points
determine parameters of a simple wax deposition model (e.g., between the stages.
Akbarzadeh and Zougari, 2008). As we mentioned in Section 1, the employed shear-removal
term is, probably, oversimplified. However, an application of a
3.2. Shear-removal term different term should be justified by experimental data, which are
currently unavailable. For example, let us compare a performance
A simple analysis of Eqs. (21) and (22) shows that if the shear- of the employed shear-removal model with the constant term
removal term is zero the porosity always increases within a certain mentioned in Section 2.1. It is clear that if the parameters of both
region adjacent to the upper layer boundary. This observation is in the shear-removal terms are identified from relatively short
line with our assumption stating that the porosity at the layer experiments of the same duration, then the constant term will
surface is maximum, which provides the layer integrity (the lower cause significantly smaller shear-removal rates at large time scales.
the porosity, the higher the deposit strength). Zero shear-removal However, a superiority of any of these shear-removal terms has
means there is no restriction on a porosity increase. The shear- not been experimentally proven. We would like to also note that
removal term is never zero in practical cases. Because the initial the strength of the deposit material increases with time due to
porosity at the surface is assumed to be maximal for a given set of aging. It is clear that the shear-removal rate should decrease with
flow conditions, it can either decrease or remain constant as the a decrease in the deposit porosity but a corresponding depen-
deposit layer grows, assuming the shear stress applied to the layer dence could be determined only experimentally and is currently
surface remains the same with time (the shear stress is nearly unknown.
constant if the layer is thin). A simple numerical analysis shows
that the conventional shear-removal term (κΘ in Eq. (22)) is not
applicable to modeling a relatively short initial stage of the deposit 3.3. Using Taylor–Couette device for identification of the
layer formation. On the one hand, when the deposit layer is very model parameters
thin, extremely high values of the shear-removal coefficient κ are
needed to prevent a porosity increase at the deposit surface. A flow loop-based experimental technique requires relatively
On the other hand, a lower coefficient κ is necessary to adequately high volumes of fluids and high flow rates to imitate flow
describe the deposit layer thickness evolution over a significant conditions, similar to those in the field. It was recently demon-
time period (days, months). This seeming difficulty in modeling of strated that a Taylor–Couette device is a convenient tool for wax
the shear-removal can be overcome as follows. The process of the deposition analysis (e.g., Akbarzadeh and Zougari, 2008). Both a
layer growth can be considered consisting of two stages, and high flow rate pump and a high fluid volume are not needed in
different shear-removal mechanisms are applied to each stage. this case. A volume of the laboratory Couette device, used by
At the first stage, the porosity at the layer surface remains constant Akbarzadeh and Zougari (2008), was only 150 cm3. In this device,
that means the shear stress at the deposit surface is critical; shown in Fig. 2, the inner cylinder rotates while the outer one is
i.e., the shear-removal prevents forming a structure of the higher immobile. The temperature gradient needed for wax deposition on
porosity on the layer top. If we formally maintain the same the inner surface of the outer cylinder is created by cooling its
functional form of the shear-removal term for the entire deposi- outer surface and by heating a fluid in the course of heating the
tion process duration (κΘ), the shear-removal coefficient κ at the inner cylinder. Thus, the experimental system allows maintaining
first stage gradually decreases as the layer thickness increases. constant both the fluid temperature inside a Couette device and
To calculate this coefficient, Eq. (22) is substituted into Eq. (21) and the temperature of the outer cylinder surface.

Multi-Point High Pressure


R Thermocouples Flow Control Valves

r0
Shear Cell

Mechanical
Mounting
Stand

Variable Speed
DC Motor

Fig. 2. Diagram of the Couette cell for deposition experiments.


316 D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319

To start an experiment, the cell (Couette device) is filled with oil at slightly higher than WAT to prevent wax crystal precipitation and
a pressure of interest and a temperature higher than the wax aggregation, is continuously supplied with a low flow rate to the
appearance temperature. To provide the wax deposition, the fluid Couette device, and then directed into a collection bottle.
and wall temperatures are set to the predefined values (the wall
temperature is lower than the fluid temperature). The spindle angular
4. Computational examples and discussion
velocity is set to match the shear stress at the outer wall to the wall
shear stress expected in the operator's flow line. Then, the test is run
The systematic experimental data for sufficiently long deposi-
for a chosen duration. After completion of the test, the deposit formed
tion cases needed for detailed application of the developed
on the cell wall is removed with hot toluene. After the solvent is
technique are not currently available. Nevertheless, the five para-
evaporated, the residue is weighed. The mass of this residue repre-
meters of our model are sufficient to match any consisted experi-
sents the total mass of the deposit including the entrapped oil. Then,
mental data on a laboratory scale; therefore, we do not present
the residue is analyzed to determine the amount of actual solid
simulation results vs. data obtained for not sufficiently long
deposit and the amount of occluded oil.
experiments. The target of this work is to demonstrate that the
Note that the shear stress at the Couette device wall for a turbulent
model behavior is robust and show how the model parameters
flow (Re413,000) can be calculated by the engineering model of
affect the deposition. For this purpose, we will employ a hydro-
Eskin (2010). The Prandtl–Karman equation (Schlichting and Gersten,
carbon fluid possessing certain properties and denote it as Oil A.
2000) can be employed for calculating the shear stress at the
Initially, we estimated model parameters assuming them to be
pipe wall.
similar to published experimental data (Akbarzadeh and Zougari,
As mentioned previously, the deposition model has five para-
2008). A Couette device had the following dimensions: inner
meters: kmd, φ(0,0) and φ(0,1), n and ksr. If the Couette device is
cylinder radius r0 ¼0.014 m, outer cylinder radius R¼ 0.028 m
employed, the molecular diffusion constant kmd can be determined
and height H¼ 0.07 m. The temperature inside the device was
from the static experiment. For identifying the 4 remaining para-
Tb ¼30 1C and the wall temperature Tw was 27 1C. The oil proper-
meters, the experimental device has to be run for different
ties were defined at the mean temperature Tm ¼(Tb+Tw)/2 ¼
durations in a regime providing shear stress at the wall that is
28.5 1C. The oil viscosity m at this temperature was 7 mPa s, the
equal to that in a modeled pipeline under field conditions. In the
oil density ρf was 843 kg/m3, the wax solids density ρwax was
case of using the Couette device, the inner cylinder rotation speed
900 kg/m3, the oil and wax solids heat capacity cp was 2020 J/kg K,
has to be selected to provide the needed shear stress at the outer
and the relatively low heat conductivity was selected (λ¼ 0.1 W/mK).
wall (e.g., Eskin et al., 2011). Note that Akbarzadeh et al. (2010)
The equilibrium wax concentration cm was 90 kg/m3, and the deri-
demonstrated how a wax deposition model can be tuned with
vative of the wax molecular concentration with respect to the
experimental data obtained in the Couette device.
temperature dc/dT was 0.027 kg/(m3 K). The heat transfer coefficient
To avoid depleting a hydrocarbon fluid of wax molecules, it is
between the bulk fluid and the Couette device wall was calculated as
recommended to use a flow-through modification of the Couette
(Holman, 1990)
device. Eskin et al. (2011) employed such a device for studying
asphaltene deposition. For wax deposition experiments, a hydro- 1
κbw ¼ ð25Þ
carbon fluid, maintained in a feeding bottle under a temperature ð1=kbs Þ þ ðδ=λÞ

25 30
Specific Deposit Mass, g/m2

25
20
n=1.5,δ0=0.1 μm
20
15
δ, μm

15
10 n=1.5, δ0=0.1μm
10
5 5

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time, h Time, h

0.8 1
0.7 0.9
0.6 0.8
n=1.5, δ0=0.1 μm
0.5 n=1.5, δ0=0.1 μm 0.7
0.4 0.6
m
φ

0.3 0.5
z

0.2 0.4
0.1 0.3
0 0.2
0 5 10 15 20 25 30
0.1
Time, h
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
φ

Fig. 3. (a) Specific deposit mass vs. time at different shear-removal rates. (b) Deposit layer thickness vs. time at different shear-removal rates. (c) Mean deposit layer porosity
vs. time at different shear-removal rates. (d) Porosity distributions across normalized deposit layer thickness for different time moments at different shear-removal rates.
D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319 317

where kbs is the coefficient of heat transfer between the bulk fluid and different computations (n¼1.5 and 2.5). The shear-removal factor was
the deposit layer surface. set to be ksr ¼3  10−7 (Pa s)−1. In Fig. 4a one can see how the deposit
The coefficient kbs was determined by an empirical correlation mass changes with time. For the larger exponent n the deposit mass
for a turbulent Couette flow as (Ball et al., 1989) increases slower. This result can be explained as follows. The larger the
exponent n, the faster the diffusivity decreases with a decrease in the
14:54 þ 0:74Re0:53
κ bs ¼ λf ð26Þ porosity. Thus, for the larger exponent n, the porosity is distributed
2ðR−r 0 Þ
more uniformly across the layer (see Fig. 4d) and a larger mass of solid
where Re ¼ ωr 0 ðR−r 0 Þ=ν, ω is the angular velocity. material is removed by shear. In Fig. 4b we present the layer thickness
The spindle rotation speed was assumed to be 60 Hz for all the vs. time. The thickness increases faster for the higher exponent n
numerical examples. values, because the porosity reduction across the layer is slower in this
The molecular diffusion constant was assumed equal to kmd ¼ 3 case. This phenomenon is also illustrated in Fig. 4c where an evolution
.7  10−10 that yields the wax molecular diffusivity D¼5.2  10−9 m2/s, of the mean porosity with time is shown.
which is within the range reported in different studies (D∼10−10– For further numerical examples we significantly increased the
10−8 m2/s according to Gudmundsson (2010)). To minimize the time, initial layer thickness (δ0 ¼15 μm). This value is comparable with
during which the porosity distribution changes significantly, most of heights of asperities forming pipe wall roughness. We also noticeably
calculations were performed at a very small initial deposit thickness reduced the shear-removal factor (ksr ¼ 3  10−8 (Pa s)−1). The calcula-
(δ0 ¼0.1 μm). In Fig. 3(a–d) we showed the results obtained at the tions were performed for the different diffusivity exponents n¼2.5
different shear-removal factors ksr ¼ 5  10−7 and ksr ¼3  10−7 (Pa s)−1. and 3.5, and the different initial porosity distributions of φ(0)¼0.74,
The diffusivity exponent was selected to be n¼1.5 that is a relatively φ(1)¼ 0.76 and φ(0)¼0.89, φ(1)¼ 0.91. The initial mean porosities are
low value for a porous material. The initial porosity distribution was φm0 ¼0.75 and φm0 ¼0.9. It is important to emphasize that for these
assumed to be linear and nearly uniform: φ(0)¼0.74 and φ(1)¼ 0.76. set of computations we introduced the different and higher thermal
In Fig. 3a one can see the specific mass accumulated on the outer wall conductivities of oil and solid wax (λf∼0.14 and λwax∼0.2 W/m K).
of a Couette device as a function of time. Obviously, the larger the According to the idea, formulated in Section 2.1, the deposit thermal
shear-removal factor is, the smaller the deposit mass. In Fig. 3b we conductivity λ was assumed to be a function of the mean deposit
showed how the deposit layer thickness increases with time, and in porosity φm. Therefore, at each computation time step, the deposit
Fig. 3c one can see the mean porosity evolution. The mean porosity thermal conductivity was calculated as follows:
decrease is faster for the higher shear-removal rate, because the higher
1
the shear-removal rate, the more intensely the material with the λðφm Þ ¼ R 1 ð27Þ
highest porosity at the layer top is removed, which leads to a faster 0 ð1=λðφðzÞÞÞdz

decrease in the mean porosity level. In Fig. 3d we showed the porosity The function λ(φ) was calculated by the Maxwell's theoretical
distributions across the layer at the different time moments. Because equation (e.g., Singh et al., 2000):
the initial deposit thickness is small, the diffusion is able to cause a
ω þ 2−2ð1−φÞð1−ωÞ
rapid change in the porosity distribution; therefore, the porosity level λ ¼ λf ð28Þ
ω þ 2 þ ð1−φÞð1−ωÞ
becomes low after a relatively short time (t¼23.6 h). The second set of
figures (Fig. 4(a–d)) illustrates the diffusivity effect on deposit layer where ω¼ λwax/λf.
formation. The results were obtained for the same parameters of Eq. (28) was derived for a system composed of a fluid and
the system, except the diffusivity exponents were different for the spherical particles of low concentration. Singh et al. (2000, 2001)

25 30
n=1.5
Specific Deposit Mass, g/m2

n=1.5
n=2.5
20 25 n=2.5

ksr=3 10−7 1/Pa s, δ0=0.1 μm


20
15
δ, μm

15 k sr =3 10−7 1/Pa s
10 δ0=0.1 μm
10
5 5

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Time, h Time, h

0.8 1
0.7 n=1.5
0.9
n=2.5
0.6 0.8 −7
k sr=3 10 1/Pa s
−7
ksr=3 10 1/Pa s δ 0 =0.1 μm
0.5 δ 0=0.1 μm 0.7
0.4
φm

0.6
0.3 0.5
z

0.2 0.4
0.1 0.3 t=0.0 h
t=2.0 h, n=1.5
0 0.2 t=10.0 h, n=1.5
t= 23.6 h, n=1.5
0 5 10 15 20 25 30 t= 2.0 h, n=2.5
0.1 t= 10.0 h, n=2.5
Time, h t= 23.6 h, n=2.5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
φ

Fig. 4. (a) Specific deposit mass vs. time at different diffusivity exponents. (b) Deposit layer thickness vs. time at different diffusivity exponents. (c) Mean deposit layer
porosity vs. time at different diffusivity exponents. (d) Porosity distributions across normalized deposit layer thickness for different time moments at different diffusivity
exponents.
318 D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319

and others employed this equation for modeling, although its affect the deposition process. We compared the computation results,
applicability to the wax deposit is questionable. shown in Fig. 5(a–d) with those obtained for the constant deposit layer
The computation results are shown in Fig. 5(a–d). In Fig. 5a one can conductivity λ¼ 0.1 W/m K and found out that using the different oil
see the deposit mass vs. time. As it is already illustrated in Fig. 4a the and wax conductivities caused only small increases in the deposit
deposit mass increases faster for the smaller diffusivity exponent. Note mass and the deposit thickness. The maximum increase did not
also that due to the significantly lower shear-removal rate compared exceed 10%. Moreover, it turned out that the distributions of the mean
to the previous calculations, the larger deposit masses are accumulated layer porosity with time and the distributions of the porosity across
on the wall. However, the slopes of the curves are significantly smaller the deposit layer at the different time moments are nearly the same
compared to the cases with the smaller initial deposit thickness. The for the different thermal conductivities employed.
thicker the layer, the slower the reduction of the porosity level is, and Thus, the presented numerical examples demonstrate a wide
therefore, the larger the time of slowing down the deposit growth. range of possible deposit layer growth scenarios for the various
In Fig. 5b we showed the deposit layer thicknesses vs. time. It is model parameters. As we mentioned previously, fitting computa-
interesting to note that in the case of the larger diffusivity exponent tional results to experimental data for wax deposition on a
(the smaller diffusivity through porous media), the deposit thickness laboratory scale is the easy task. In almost any of the published
initially grows faster but after a certain period of time the thicknesses works an excellent match of calculated deposit layer thickness (or
become almost equal. This observation is primarily caused by the the total deposit mass) to corresponding experimental data was
lower initial porosity in the 2nd case characterized by the lower demonstrated. However, such data fit does not guarantee a proper
diffusivity. On the one hand, the higher the porosity, the higher the deposition process scaling to field conditions, including realistic
deposit thickness growth rate (Eq. (22)). On the other hand, the larger time scales. Through the computational examples we demon-
the thickness, the larger the shear-removal rate (see Eq. (10)), and the strated how strongly the model parameters affect the deposit layer
lower the diffusivity, the slower the deposit mass accumulation structure and dynamics. Thus, an adequate deposition modeling
(Fig. 5a). Thus, these counteracting factors lead to the nearly equal requires a proper description of the deposit layer formation. The
layer thicknesses (Fig. 5b). In Fig. 5c one can see how the mean model that we suggested can be successfully used for accurate wax
porosities change with time, and in Fig. 5d we show the porosity deposition modeling if the model parameters are carefully identi-
distributions at different time moments for both the calculated cases. fied from the lab data, because it is clear that the model takes into
It is important to note that the initially linear porosity distributions account the major deposition phenomena. Moreover, there is a
remain nearly linear during the entire deposit layer formation process sizable chance that the initial deposit layer thickness δ0 and the
in spite of a significant reduction in the porosity level. This linearity diffusivity exponent n do not significantly vary for different
can be employed for efficient engineering calculations of the deposit deposition cases and will be fixed after these parameters are
aging in long oil transport pipelines, which are beyond the scope of reliably identified. Then, in contrast to simpler deposition models
the present work. that can be found in literature, our approach additionally requires
We would like to also emphasize that setting the different and only the initial porosity distribution. For example, instead of the
higher thermal conductivities of fluid and solid wax in this case (0.14 initial mean porosity needed for the oversimplified aging model
and 0.2 vs. 0.1 W/m K) along with the recalculation of the mean (Singh et al., 2000; Hernandez et al. 2003), our model requires the
deposit layer thermal conductivity at each time step, did not strongly initial porosity distribution; i.e., φ(0) and φ(1).

140 200
Specific Deposit Mass, g/m2

120 0
150
100

80
δ, μm

100
60

40
50
20

0 0
0 100 200 300 400 0 100 200 300 400
Time, h Time, h

1 1
0.9 0.9

0.8 0.8
0.7
0.7
0.6
m

0.6
φ

0.5
z

0.5
0.4
0.4
0.3
0.3 0.2 δ μ

0.2 0.1
0 100 200 300 400
0
Time, h 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
φ

Fig. 5. (a) Specific deposit mass vs. time at different diffusivity exponents. (b) Deposit layer thickness vs. time at different diffusivity exponents. (c) Mean deposit layer
porosity vs. time at different diffusivity exponents. (d) Porosity distributions across normalized deposit layer thickness for different time moments at different diffusivity
exponents.
D. Eskin et al. / Chemical Engineering Science 97 (2013) 311–319 319

5. Conclusions Subscripts

A model of the wax deposit layer formation on the wall has d deposit or deposition
been developed. The deposit layer was assumed to be a porous f fluid
layer, whose growth is accompanied with a porosity reduction. m mean
The porosity decrease is caused by the growth of wax crystals mov moving
composing a porous medium in a wax saturated hydrocarbon o oil
fluid. The two major model equations are an equation of diffusion p pipe
of wax molecules through a porous medium of varying porosity, sr shear-removal
and an equation for the deposit layer thickness evolution. The wax wax solids
model takes into account a diffusivity reduction with a porosity 0 initial
decrease and a shear-removal effect. The model developed has
been illustrated by computations of the wax deposition in a
Taylor–Couette device. The calculations demonstrated that the
aging rate rapidly increases with a decrease in the parameter Acknowledgments
accounting for dependence of the wax diffusivity on the layer
porosity and/or an increase in the shear-removal factor. It has been The authors are grateful to Prof. J.R.A. Pearson (Schlumberger
also shown that at the realistic initial deposit thickness, an initially Gould Research Center) for his valuable advices and discussion of
linear porosity distribution across the deposit layer nearly main- this work.
tains linearity during the entire deposition process.
References

Akbarzadeh, K., Ratulowski, J., Eskin, D., Davies, T., 2010. The importance of wax
deposition measurements in the simulation and design of subsea pipelines. SPE
Nomenclature Projects, Facilities Constr. 5 (2), 49–57.
Akbarzadeh, K., Zougari, M., 2008. A new approach in single-phase wax deposition
modeling; Part 1: Taylor–Couette system. Ind. Eng. Chem. Res. 47 (3), 953–963.
D molecular diffusivity of wax (m2/s) Ball, K.S., Farouk, B., Dixit, V.C., 1989. An experimental study of heat transfer in a
c wax mass concentration (kg/m3) vertical annulus with a rotating inner cylinder. Int. J. Heat Mass Transfer 32 (8),
cp heat capacity (J/kg/K) 1517–1527.
Burger, E.D., Perkins, T.K., Striegler, J.H., 1981. Studies of wax deposition in the trans
n exponent determining the wax diffusivity through Alaska pipeline. J. Pet. Technol. 33 (6), 1075–1086.
porous media Carson, J.K., Lovatt, S.J., Tanner, D.J., Cleland, A.C., 2005. Thermal conductivity
kbs the heat transfer coefficient between the bulk fluid and bounds for isotropic, porous materials. Int. J. Heat Mass Transfer 48, 2150–2158.
Cussler, E.L., Hughes, S.E., Ward, W.J., Aris, R., 1988. Barrier membranes. J. Membr.
the deposit layer surface (W/m2/K) Sci. 38 (2), 161–174.
kbw the heat transfer coefficient between the bulk fluid and Elam, S.K., Tokura, I., Saito, K., Altenkirch, R.A., 1989. Thermal conductivity of crude
the Taylor–Couette device wall (W/m2/K) oils. Exp. Thermal Fluid Sci. 2, 1–6.
Eskin, D., Ratulowski, J., Akbarzadeh, K., Lindvig, T., 2009. An approach to prediction
ksr shear-removal factor (Pa−1/s)
of wax and asphaltene deposition in a pipeline based on Couette device
qd deposition rate (kg/m2/s) experimental data. Comput. Methods Multiphase Flow V. WIT Trans. Eng. Sci.
R outer Taylor–Couette device radius (m) 63, 85–97.
Eskin, D., 2010. An engineering model of a developed turbulent flow in a Couette
Re Reynolds number
device. Chem. Eng. Process. 49 (2), 219–224.
r0 inner Taylor–Couette device radius (m) Eskin, D., Ratulowski, J., Akbarzadeh, K., Pan, S., 2011. Modeling asphaltene
T temperature (K) deposition in turbulent pipeline flows. Can. J. Chem. Eng. 89 (3), 421–441.
t time (s) Gudmundsson, J.S., 2010. Flow Assurance, Solids in Oil and Gas Production.
The Report. Department of Petroleum Engineering and Applied Geophysics,
U superficial flow velocity through a pipe (m/s) Norwegian University of Science and Technology, Trondheim.
z dimensionless coordinate Hayduk, W., Minhas, B.S., 1982. Correlations for prediction of molecular diffusivities
in liquids. Can. J. Chem. Eng. 60 (2), 295–299.
Hernandez, O.C., Hensley, H., Sarica, C., Brill, J.P., Volk, M., Delle-case, E., 2003.
Greek letters Improvements in single-phase paraffin deposition modeling. In: Proceedings of
SPE Annual Technical Conference and Exhibition, Denver, SPE 84502.
Holman, J.P., 1990. Heat Transfer, 7th edition McGraw-Hill, New York.
δ deposit layer thickness (m) Huang, Z., Lee, H.S., Senra, M., Fogler, H.S., 2011. A fundamental model of wax
ε shear-removal parameter (kg/m3/s) deposition in subsea oil pipelines. AICHE J. 57 (11), 2955–2964.
φ porosity Kern, D.Q., Seaton, R.E., 1959. A theoretical analysis of thermal surface fouling. Br.
Chem. Eng. 4, 258–262.
κ shear-removal constant (s−1) Lee, H.S., 2007. Computational and Rheological Study of Wax Deposition and
κt shear-removal constant at the second process stage (s−1) Gelation in Subsea Pipelines (Ph.D. Thesis). University of Michigan.
λ thermal conductivity (W/m/K) Peng, S., Hu, Q., Hamamoto, S., 2012. Diffusivity of rocks: Gas diffusion measure-
ments and correlation to porosity and pore size distribution. Water Resour. Res.
ν kinematic viscosity (m2/s)
48 (2), 1–9.
ρ density (kg/m3) Schlichting, H., Gersten, K., 2000. Boundary-Layer Theory. Springer-Verlag.
Θ characteristic time (s) Singh, P., Venkatesan, R., Fogler, H.S., 2000. Formation and aging of incipient thin
film wax–oil gels. AICHE J. 46 (5), 1059–1074.
τ dimensionless time
Singh, P., Venkatesan, R., Fogler, H.S., 2001. Morphological evolution of thick wax
τw shear stress at the wall (Pa) deposits during aging. AICHE J. 47 (1), 6–18.
υ dimensionless parameter accounting for wax properties Svendsen, J.A., 1993. Mathematical modeling of wax deposition in oil pipeline
ω angular velocity of the Taylor–Couette device spindle systems. AIChE J. 39 (8), 1377–1388.
Ukrainczyk, N., Kurajica, S., Šipušic, J., 2010. Thermophysical comparison of five
(rad/s) commercial paraffin waxes as latent heat storage materials. Chem. Biochem.
ξ axial coordinate along pipe axis (m) Eng. 24 (2), 129–137.

Das könnte Ihnen auch gefallen