You are on page 1of 10

Heavy oil transportation as a slurry

A Brito, R Cabello, L Mendoza


PDVSA Intevep, Venezuela
H Salazar
UDO, Venezuela
J Trujillo
MSi Kenny, Australia

ABSTRACT

Heavy oil transportation is especially difficult due to the high viscosities of the oil and
the complexity of the multiphase flow involved. Asphaltenes and resins have the highest
molecular weight of crude oil. This work investigates an experimental behavior of a
potential transport method for heavy oils based on slurry formed by solids of asphaltenes
and resins reincorporated into the deasphalted oil. Dynamic tests were conducted in a 1”
horizontal pipeline, with low solid contents in a laminar flow regimen. Of three models
that were evaluated to predict pressure drop in the pipeline, the Darcy model provided the
best prediction.

1. INTRODUCTION

Transportation of heavy oil from wellhead to central processing facilities (CPF) for
degassing and dehydration and from CPF to downstream facilities for upgrading or
refining purposes is a technical challenge due to the high viscosity. The continuous
increase in the world’s oil demand is leading to improvements to existing transportation
solutions or development of new ones.

Venezuela has 296.500 million barrels of heavy and extra heavy oil certified reserves,
located in the Orinoco Oil Belt. The main strategy to develop the Orinoco Oil Belt is a
combination of dilution and upgrading of the extra heavy oil. For the developments to be
feasible, however, alternative transportation solutions must be explored, such as the
addition of polar compounds to improve dilution, heating, oil-in-water emulsion, core
annular flow, foam flow and slurry transportation. [1, 6]

Virgin heavy and extra heavy oils from the Orinoco Oil Belt exhibit viscosities in the
order of several thousand of centipoises at typical surface operating conditions. This fact
imposes a significant technical challenge to develop these vast resources. By diluting
heavy oils with lighter oils or naphtha, viscosity is effectively reduced to typical ranges
of 500 to 1000cP. However, required diluent/heavy oil ratios are around 30:70. When
using core annular flow or oil in water emulsions, typical water to oil ratios required are
also around 30:70. For these transportation methods, addition of diluent or water means a
larger pipe size, with higher CAPEX, besides other technical and economic significant

© BHR Group 2013 Multiphase 16 95


challenges. A sustainable source of diluent means an approach of vertical integration
with an upgrader as the core of the strategy (high CAPEX), or securing long-term supply
contracts to acquire diluent from reliable third parties (high OPEX). A long-term source
of water and further proper treatment and disposal of that water impose important
environmental and investment issues to address. [1]

Sanier et al. (2004) studied a potential heavy oil transportation method based on a solid-
liquid dispersion (slurry), where asphaltenes precipitated using saturates of low
molecular weight are reincorporated into the deasphalted oil (DAO) to obtain a slurry or
suspension of non-colloidal particles of relatively low viscosity that fluidizes the oil. One
of the main findings from Sanier et al. (2004) work are: a) the slurry exhibits
approximately 75% less viscosity than the original crude oil, and b) precipitated
asphaltenes become progressively colloidal with time. As expected, asphaltenes kinetics
are relatively slow. Evolution from solid to colloidal state can be accelerated by
increasing temperature. Also, Saniere et al. (2004) found there is a critical asphaltenes
concentration (around 10%): under this critical concentration the asphaltenes behave as
colloidal particles dispersed in oil and above this value the particles of asphaltenes
overlap and the viscosity increases dramatically [2-5].

Feasibility of any deasphalting process is highly dependent on a reliable source of low


molecular weight n-alkanes, which are well known for inducing precipitation of
asphaltenes [9]. As an example, typical n-heptane to oil ratios required to precipitate
asphaltenes in Venezuelan heavy oils are around 7:1. This high requirement of low
molecular weight n-alkanes imposes a technical challenge to any solvent deasphalting
process: a highly efficient recovery process to keep n-alkanes make up rates as low as
possible.

A proprietary solvent deasphalting process under non-severe pressure and temperature


conditions produces a DAO with excellent transport properties. Measured viscosities of
DAO used in this study range from 0.0385 to 0.0464 Pa.s, at least three orders of
magnitude lower than viscosity of virgin heavy oil. With addition of up to 12% in weight
of solid residue coming from that deasphalting process, a solid in liquid suspension or
slurry is created. Such slurry exhibits mixture viscosities from 0.0473 to 0.1237 Pa.s,
which are, as expected, higher than DAO viscosity, but still significantly lower than the
original heavy oil viscosity. Resulting mixture viscosity is comparable to typical
viscosity of heavy oil diluted with naphtha, but without the need of a proportional
increase of flow rate caused by the addition of diluent. Same argument applies when
compared to a CAF and oil in water emulsion transport methods. This translates into
lower requirements of pipe line size, and therefore in CAPEX.

This paper presents the results from an experimental test to evaluate hydrodynamic
properties of solid – liquid suspensions or slurries, prepared with DAO (also known as
maltenes) as the liquid phase and the asphaltic residue as the solid phase, both phases
obtained from a proprietary de-asphalting process at relatively low pressure and
temperature. The link between the morphology of these slurries and their rheology is
explained in more detail by Brito et al. [6]. The slurry hydrodynamic properties are
evaluated using a flow loop of Ø0.025m and slurry properties like density, viscosity and
morphology were studied at the beginning and completion of each test and 24 hours later.
Also, three different models were evaluated to predict the frictional pressure drop in a
horizontal pipeline: Turian & Yuan 1977, Turian et al 1987, and Darcy.

96 © BHR Group 2013 Multiphase 16


2. EXPERIMENTAL DESCRIPTION

2.1. Set-up
A new facility was designed and built at PDVSA INTEVEP to evaluate the slurry
behavior. Experiments were carried out in a test loop of Ø0.025m internal diameter (ID).
Figure 1 shows the experimental facility diagram, the total length of the loop is 6 m of
carbon steel pipeline, which is conformed by two sections: the first one is a 3.5 m length
section for flow development; the second is the test section, of 2.5 m length, equipped
with transmitters for: pressure, temperature and differential pressure. Differential
pressure transducers are used to measure the pressure drop between pressure taps.
Downstream to the test section, the dispersion returns to the tank in which the slurry is
continuously agitated. The slurry is pumped using a gear pump with a variable speed
control to regulate the flow rate, the maximum dispersion flow rate is 3.6m3/hr and it is
measured by deviating the flow to a small tank with a level transmitter.

Tank

Pump

Figure 1. Experimental test loop diagram

2.2. Fluid system


The slurry is prepared with de-asphalted oil (also known as maltenes) as the liquid phase
and the asphaltic residue as the solid phase, both phases obtained from a proprietary de-
asphalting process at relatively low pressure and temperature. Properties and SARA
composition of both the solid and liquid hydrocarbons that were mixed to prepare the
slurries are presented in Table 1. The solid phase is composed mainly by asphaltenes and
resins, and is prepared and sieved with a special procedure to keep particle size up to 500
μm, this solids are incorporated in the liquid under an appropriate mechanical stirring for
20 min.

Table 1. Properties and S.A.R.A of liquid and solid phases


Density Viscosity
Saturates Aromatics Resins Asphaltenes
Fluid @15°C @135°C
%p/p %p/p %p/p %p/p
Kg/m3 cST
De-asphalted oil 907 20.57 15 56 29 0
Residue oil 1120 - 14 15.4 31.6 39

© BHR Group 2013 Multiphase 16 97


2.3. Test matrix
Test conditions are presented in Table 2. For each experimental point the dispersion
properties are analyzed for each solid concentration and mixture velocities between 0,2 –
2,3 m/s. Samples of the slurry were taken previous to start of the dynamic test (initial
stage), just after running the dynamic test (middle stage), and 24 hours after finalizing the
dynamic test (final stage).

The dispersion morphology is analyzed using optical microscopy. The viscosity of the
dispersion is assessed using a viscometer Anton Paar model MCR 301 according to the
ASTM D 7483 standard test method. Finally, the dispersion density is quantified using a
densitometer Anton Paar DMA 4500 according to the ASTM D 4052 standard test
method.

Table 2. Test matrix for the slurry transportation test


Number of
Mixture velocities (m/s) Solid Concentration (%)
experiments
36 0.2, 0.36, 0.72, 1.26, 1.62 and 2.33 0, 1, 3, 6, 8 and 12

3. PRESSURE DROP CORRELATIONS MODELS

Three different models were evaluated to predict frictional pressure drop in a horizontal
pipeline: Turian & Yuan 1977, Turian et al 1987, and Darcy models. The Turian & Yuan
1977 and Turian et al 1987 models consider four different flow patterns: homogeneous
flow, heterogeneous flow, saltation flow and flow with a stationary bed (Figure 2). [7, 8,
10]

a) Homogeneous flow b) Heterogeneous flow

c) Saltation flow d) Flow with a stationary bed

Figure 2. Flow patterns considered in the Turian & Yuan 1977 and Turian et al
1987 models, 1977 [7]

To provide a meaningful flow regime classification criterion, Turian & Yuan 1977
proceeded to correlate two contiguous flow regimes and defined a dimensionless group
named the regime number Rab:

v2 (1)
R ab =
⋅ D ⋅ g ⋅ (s − 1)
αt βt γt
K ⋅ Cs ⋅ fl ⋅ CD

98 © BHR Group 2013 Multiphase 16


where fl is the liquid phase friction factor, Cs is the mean solid concentration, CD is the
drag coefficient for free-falling sphere and K, α, β, γ, δ are constants of the correlation
that depends on the flow pattern in the pipeline, that are presented in Table 3 and Table 4.

The transition point between flow regimes a and b is given by Rab = 1. Moreover, since v
increases, the change in fw is relatively small in the turbulent flow region. It can further
be inferred that Rab > 1 corresponds to the faster flow regime, while Rab < 1 corresponds
to the slower flow regime.

The mixture friction factor fm used by Turian & Yuan 1977 and Turian et al 1987 models
to calculate frictional pressure drop gradient is presented in the equation 2.

δ
α ⎡ v m2 ⎤ (2)
fm − fl = KCs fwβ CDγ ⎢ ⎥
⎣ Dg (s − 1) ⎦

Table 3. Parameters for friction factor calculation in Turian & Yuan (1977) model
Flow Regimen K α β γ δ
Flow with a stationary bed 0,4036 0,7389 0,7717 -0,4054 -1,096
Saltation flow 0,9857 1,018 1,046 -0,4213 -1,354
Heterogeneous flow 0,5513 0,8687 1,200 -0,1677 -0,6938
Homogeneous flow 0,8444 0,5024 1,428 0,1516 -0,3531

Table 4. Parameters for friction factor calculation in Turian et al. (1987) model
Flow Regimen K α β γ δ
Flow with a stationary bed 12,127 0,7389 0,7717 -0,4054 -1,096
Saltation flow 107,09 1,018 1,046 -0,4213 -1,354
Heterogeneous flow 30,115 0,8687 1,200 -0,1677 -0,6938
Homogeneous flow 8,538 0,5024 1,428 0,1516 -0,3531

Liquid friction factor was correlated to a liquid Reynolds number defined as:

v m Dρ L (3)
Re L =
μL

Finally, the frictional pressure gradient is determined with the Darcy equation, using
friction factor determined with equation (1) and considering the mixture density and
velocity:

ΔP ρ v2 (4)
− = 2f m m m
L D

When evaluating the slurry frictional pressure drop with a simpler Darcy model, the solid
– liquid mixture is considered as a homogeneous single phase, with a Darcy friction
factor correlated to a mixture Reynolds number defined in this study as: [10]

v m Dρ m (5)
Re m =
μm

© BHR Group 2013 Multiphase 16 99


4. RESULTS AND DISCUSSION

The morphology of the slurry was analyzed using optical microscopy. Results show that
samples are not colloidal; the slurry is conformed by detectable solid particles with size
over 1μm and irregular shape. Slurry samples were taken previous to start of the
dynamic test (initial stage), just after running the dynamic test (middle stage), and 24
hours after finalizing the dynamic test. The results show that the precipitated asphaltenes
become progressively colloidal with time (Figure 3) and solid concentration (Figure 4).
Viscosity increase with time confirms such behavior (Figure 5).

Figure 3. Morphology of the slurry of 6% slurry sample,


optical microscopy, T = 20°C

Figure 4. Morphology of the slurry sample for different solid concentrations,


optical microscopy, T = 20°C

It is important to point out that the slurry maximum change of viscosity with time is
approximately 45%, occurring at the maximum solid concentration of 12%, with a final
dynamic viscosity around 0.16 Pa.s, which is at least three orders of magnitude lower
than the typical viscosity of the virgin heavy oil of the Orinoco Oil Belt [3].

When comparing results obtained during initial, middle and final stages, small
differences in viscosity are found for solids concentrations below 12%. However, for
12% of solids concentration, viscosity differences with time are clearly noticeable. This
could be explained by a re-dissolution of asphaltenes into the maltenes, returning to their
colloidal state. The 12% threshold seems to coincide with findings of Sanier et al (2004)
for the critical asphaltene concentration in slurry studied formed with Orinoco Belt heavy
oils. (Figure 5)

100 © BHR Group 2013 Multiphase 16


Figure 5. Change of viscosity with time for a 6% slurry sample, T = 20°C

Relative viscosity is defined for the scope of this work as the ratio of mixture viscosity at
certain solid concentration to the measured maltenes viscosity (i.e. DAO with no solids).
It is presented in Figure 6 as a function of solid concentration and mixture velocity. It
was observed that the viscosity of the mixture increases with the solid concentration, as
expected and for most solid concentrations, increasing mixture velocities, with
corresponding increasing shear stress, do not affect the relative viscosity. However, for
solid concentrations of 12%, as mixture velocity increases above 0.7 m/s, there is a
progressive reduction in the relative viscosity. This last phenomenon is due to the viscous
dissipation of energy inside the pump and pipes on the test loop, which promotes an
increment in temperature up to 5°C.

280%
260%
Relative viscosity (%)

240%
220%
200%
180%
160%
140%
120%
100%
12
0.2 8
0.4 6
0.7 3
1.3
1 Solids concentration (% wt.)
Mixture velocity (m/s) 1.6
2.3 0

Figure 6. Relative viscosity as a function of the solid concentration and mixture


velocity

As solid – liquid dispersions are known to exhibit a non Newtonian behavior, that could
be the first argument to explain such behavior. However, similar to the results obtained
by Saniere et al. (2004), the slurry exhibits a clear Newtonian behavior and does not have
yield strength. (Figure 7). [5]

© BHR Group 2013 Multiphase 16 101


12

2
10 R =1

Shear Stress (Pa) 8


2
R =1
6 R =1
2

2
R =1
4 2
R =1
C 0%
C 3%
2
C 6%
C 8%
C 12%
0
0 20 40 60 80 100
Shear Rate (1/s)

Figure 7. Shear stress vs. shear rate for different slurry concentrations

The flow regimen in the pipeline for the entire evaluated test matrix was laminar flow
with Reynolds number below 1400 (Figure 8). As expected, pressure gradient in the
pipeline increases with the Reynolds Number and solid concentration in the slurry. The
last is mainly due to the increment in the viscosity of the slurry with solid concentration.

12000
Experimental Pressure Gradient (Pa/m)

10000

8000

6000

4000

2000

0
0 200 400 600 800 1000 1200 1400
Reynolds Number

C 0% C 1% C 3% C 6% C 8% C 12%

Figure 8. Pressure gradient vs Reynolds Number

Finally, three different models were used to predict the pressure drop in the pipeline:
Darcy equation for frictional pressure drop considering the slurry as an homogeneous
single phase; the other two pressure drop models were Turian and Yuan (1977) and
Turian et al. (1987), which were developed for a solid-liquid phase transportation. In the
case of Turian and Yuan (1977) and Turian et al. (1987) models, it was necessary to
determine the flow pattern associated to each dynamic condition, obtaining for all the
experimental points of test matrix that the flow pattern was homogeneous flow. Knowing
the flow pattern, parameters were selected accordingly from Table 3 and Table 4 to
estimate the friction factor for Turian & Yuan and Turian et al. models, respectively.

102 © BHR Group 2013 Multiphase 16


Absolute average errors obtained with the three models when compared to experimental
data were estimated to be 11% for Darcy model, 54% for Turian and Yuan (1977) model
and 667% for Turian et al. (1987) model. The spread of data around the correlations are
shown in Figure 9. The best prediction performance was obtained using Darcy model,
with an average absolute error of approximately 11% and a standard deviation of 6%
when compared to the experimental pressure drop.

Figure 9. Pressure gradient calculated vs. experimental

SUMMARY AND CONCLUSIONS

• Slurry prepared with up to 12% of solids and DAO obtained from a proprietary
solvent deasphalting process were evaluated under dynamic conditions. Six different
solid concentrations were evaluated in an experimental flow loop at six different
mixture velocities. All experimental points exhibit Reynolds numbers <1400
(laminar flow) and homogeneous flow pattern.
• The obtained slurry viscosity is significantly lower than the viscosity of original
heavy oil (three orders of magnitude lower) and comparable to viscosities obtained
when diluting oil with naphtha at 70:30 oil/naphtha ratios, but without the additional
increase in volumetric flow rates and pipe size.
• The slurry exhibits a Newtonian behavior for all solid concentrations studied, with no
yield strength, which means no significant additional pump discharge pressure is
required for re-start operations for this particular issue. However, an increasing pump
discharge pressure can be required after a prolonged shutdown due to a viscosity
increase with time caused by progressive re-dissolution of asphaltenes.
• Slurry morphology and viscosity are time dependent, as solid phase dispersed in
DAO progressively return to their colloidal state. After 24 hours, slurry viscosity
increases up to 45% when compared to the initial stage for the maximum solids
concentration of 12%, to a maximum value of 0.16 Pa.s, which is still three orders of
magnitude lower than viscosity of original heavy oil. Asphaltenes re-dissolution is a
relatively slow phenomenon as asphaltenes kinetics is slow. However, additional
studies are required to assess long-term stability of slurry, to properly address
requirements for re-start operations after prolonged shutdowns.

© BHR Group 2013 Multiphase 16 103


• The pressure drop in the pipeline can be estimated using a simple Darcy frictional
pressure drop model, considering the suspension as a homogeneous phase, with an
absolute average error of 11%.

ACKNOWLEDGEMENT

The authors thank the support of PDVSA Intevep.

REFERENCES

1. Brito, A., Trujillo,J. Considering multiphase flow issues for selection of heavy oil
transportation methods in Venezuela.WHOC11-107-2011
2. Argillier, J., Henaut, I., Gateau, P. Method of transporting heavy crude oils in
dispersion. US Patent Application 20060118467. 2006.
3. Argillier, J., Henaut, I., Gateau, P., Heraud, J., Glenat, P. Heavy Oil Dilution.
SPE/PS-CIM/CHOA 97763. International Thermal Operations and Heavy Oil
Symposium. Alberta, Canada, 1-3 November 2005.
4. Martínez-Palou, R., Mosqueira, M., Zapata-Rendón, B., Mar-Juárez, E., Bernal-
Huicochea, C., Clavel-López, J., Aburto, J. Transportation of heavy and extra-
heavy crude oil by pipeline: A review. Jouran of Petroleum Science and
Engineering. (2010), doi: 10.1016/j.petrol.2010.11.020.
5. Saniere. A., Hénaut. I., Argillier, J. Pipeline transportation of heavy oils, a
strategic, economic and technological challenge. Rev. IFP, Vol.59, N°5, p.455-466,
2004.
6. Brito, A.; Salazar, H.; Cabello,R.; Mendoza, L.;Trujillo, J.; Alvarez, L. Heavy oil
transportation as a solid-liquid dispersion. I Congreso de la Sociedad Venezolana
de Mecanica de Fluidos, Margarita, 2012.
7. Turian, R and Yuan, T (1977). Flow of slurries in pipelines. AICHE Journal, vol
23, 232-243.
8. Turian, R; Hsu, F y Ma, T (1987). Estimation of the critical velocity in pipeline
flows of slurry. Powder technology, vol 51, 35-47.
9. Akbarzadeh, K., Hammami, A., Kharrat, A., Zhang, D., Allenson, S., Creek, J.,
Kabir, S., Jamaluddin, A., Marshall, A.G., Rodgers, R.P., Mullins, O.C.,
Solbakken, T. Asphaltenes—Problematic but Rich in Potential. Oil Review.
Summer 2007. Schlumberger.
10. Wallis, G. One dimensional two phase flow. McGrawn-Hill.1969

104 © BHR Group 2013 Multiphase 16

You might also like