Anda di halaman 1dari 39

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

net/publication/245332211

The Asphaltene and Wax Deposition Envelopes

Article  in  Fuel Science and Technology International · January 1996


DOI: 10.1080/08843759608947560

CITATIONS READS

47 453

1 author:

Kosta Leontaritis
AsphWax Inc
40 PUBLICATIONS   903 CITATIONS   

SEE PROFILE

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

Simulation of Near Wellbore Asphaltene-Induced Formation Damage View project

All content following this page was uploaded by Kosta Leontaritis on 21 October 2017.

The user has requested enhancement of the downloaded file.


The Asphaltene and Wax Deposition Envelopes

by

Kosta J. Leontaritis, Ph.D.


President, AsphWax Inc.
5406 David's Bend Drive
Sugar Land, Texas 77479

To be presented at the Symposium on


THERMODYNAMICS OF HEAVY OILS AND ASPHALTENES
1995 AIChE Spring National Meeting
March 19-23, 1995
Houston, Texas

January, 1995
2

The Asphaltene and Wax Deposition Envelopes


Kosta J. Leontaritis, Ph.D., President, AsphWax Inc.
5406 David's Bend Drive, Sugar Land, Texas 77479

Abstract
Asphaltene and wax phase behavior is quite different than the conventional "PVT" phase
behavior. Asphaltenes exhibit a behavior at some thermodynamic states called
flocculation. That is, asphaltene particles or micelles aggregate or flocculate into larger
aggregates or flocs. The locus of all thermodynamic points in a P-T-x phase diagram at
which flocculation occurs is called the Asphaltene Deposition Envelope (ADE). Paraffin
waxes, on the other hand, at some thermodynamic states, exhibit the phenomenon of
crystallization. The locus of all thermodynamic points in a P-T-x phase diagram at which
wax crystallization occurs is called the Wax Deposition Envelope (WDE). Asphaltene
flocculation can be both reversible and irreversible (as expected from the asphaltene
colloidal nature). Wax crystallization is generally a reversible process. However,
paraffin waxes more than often precipitate together with resins and asphaltenes (which
are said to be responsible for the observed irreversible thermodynamic phenomena).
Hence, some wax precipitation is occasionally reported as irreversible. Measurement of
the ADE and WDE boundaries, at in-situ conditions, is a challenging task. Measurement
of asphaltene and wax solubility at in-situ conditions inside the ADE and WDE is even
more challenging. The ADE and WDE data have practical significance and are very
useful for modeling of reservoir fluid behavior. Asphaltene and wax data for a number of
oils are presented and analyzed. This paper introduces and compares two powerful
thermodynamic diagrams that summarize the phase behavior of asphaltenes and waxes,
the ADE and WDE.

Definitions
For the purposes of this paper we will adhere to the following definitions:
Asphaltenes:
Are highly condensed polyaromatic structures or molecules, containing heteroatoms (i.e.,
S, O, N) and metals (e.g., Va, Ni), that exist in petroleum in an aggregated state in the
form of suspension and are surrounded and stabilized by resins (i.e., peptizing agents).
They are known to carry an electrical charge, and thought to be polydisperse.
Asphaltenes are a solubility class, hence, they are not pure, identical molecules. Pentane
and Heptane are the two most frequently used solvents for separating asphaltenes from
crude oil. The prefix n-Pentane or n-Heptane asphaltenes refers to the solvent used for
their separation. The composition of n-Pentane asphaltenes is different from that of n-
Heptane asphaltenes.
Resins:
Are aromatic and polar molecules, also often containing heteroatoms and metals, that
surround the asphaltene structures and are dissolved in the oil and help keep the
asphaltenes in suspension. They are surface active and, at some thermodynamic states,
form their own reversible micelles. They are polydisperse and have a range of polarity
and aromaticity. Resins are considered to be pre-cursors to asphaltenes.
Paraffin Waxes:
3

Primarily aliphatic hydrocarbons (both straight and branched chain) that change state
from liquid to solid during conventional oil production and processing operations. In
addition to aliphatics, field deposits usually contain aromatic, naphthenic, resin, and
asphaltenic molecules as well. The combined mass is called wax. Paraffin waxes usually
melt at about 110°-160° F. Field waxes contain molecules that can have melting points in
excess of 200° F.
Asphalt:
… is the residual (non-distillable) fraction of crude oil that contains suspended
asphaltenes, resins, and the heaviest aromatic and paraffinic components of oils. Propane
has been traditionally a very efficient and convenient solvent for separating asphalt from
petroleum. However, the latest commercial processes use other more efficient solvents
for asphalt separation.

Background
Asphaltenes
A review of all asphaltene literature is beyond the scope of this paper. However, a
minimum review leading into the basic theory of the paper will be presented. The
conventional equilibrium thermodynamic approaches for measuring and predicting the
phase behavior and PVT properties of hydrocarbon fluids are based on the assumption
that the molecules comprising these fluids are randomly and chaotically mixed in
molecular state within a given phase. For fluids that do not contain substantial amounts
of resins and asphaltenes, this assumption is reasonably accurate; hence, molecular
equilibrium thermodynamic approaches are applicable and generally successful in
predicting fluid phase behavior. If a hydrocarbon fluid contains substantial amounts of
asphaltenes and resins, then a portion of these molecules usually exists in an aggregated
state forming a so-called colloidal suspension. The colloidally suspended molecules form
a separate particulate phase which may be stable and remain suspended indefinitely or
may flocculate and form larger particles which settle at the conditions of the system and
drop out. Conventional molecular equilibrium thermodynamics does not deal with this
type of colloidal phase behavior and, as a result, new, more suitable measurement and
modeling techniques are required.

Pfeiffer and Saal (1940) gave the following description for the existence of asphaltenes
and resins in bitumens, "...In the asphaltic bitumens, the asphaltenes are centers of
micelles which are formed by adsorption, and perhaps partly by absorption, of part of the
maltenes on the surfaces or in the interiors of asphaltene particles. When the entire
system contains sufficient constituents for the formation of the outer region of the
micelles, the asphaltenes are fully peptized and able to move through the bitumen as
freely as the viscosity of the intermicellar phase permits...If, however, there is a shortage
of asphaltic resins, part of the forces causing the formation of the micelle are not
compensated by adsorption of asphaltic resins and the micelles will be subjected to
mutual attraction."

The picture that Pfeifer and Saal described then for bitumens is essentially the same
depicted in Fig. 1 for asphaltenic crude oils. Two years later, Swanson (1942)
demonstrated clearly for the asphalts he studied that resins do indeed play a peptizing role
for the asphaltene aggregates. Furthermore, he demonstrated that there is a critical resin
4

concentration in the maltene or continuous phase below which asphaltenes agglomerate


or flocculate and above which they remain in stable suspension. More recently,
Murzakov et al. (1980) using gravimetric sedimentation experiments halso demonstrated
the beneficial role of petroleum resins on the colloidal stability of asphaltene-containing
disperse systems.

The concept of the critical resin concentration described above was very crucial in the
development of the theory of the Thermodynamic-Colloidal Model (Leontaritis and
Mansoori, 1987).

Koots and Speight (1975) demonstrated more convincingly the relation of resins to the
stability of asphaltenes. They stated that their results indicated that petroleum asphaltenes
are not soluble in their corresponding oil fractions. Nor are asphaltenes from one crude
soluble in another crude. It was, however, possible to bring about peptization of the
asphaltenes by addition of the corresponding resins. Moschopedis and Speight 18 studied
the role played by the oxygen functions in the hydrogen bonding interactions which occur
between the asphaltene and resin entities of Athabasca bitumen. Their results showed
that hydrogen bonding occurs readily between these fractions and is a feasible mechanism
by which the asphaltenes are peptized by resins.

Speight et al. (1985) gave a critical review of the molecular weight and association of
asphaltenes. They reviewed the different methods (i.e. VPO, size exclusion
chromatography, ultrafiltration, ultracentrifugation, viscosity, small angle X-ray
scattering, infrared spectroscopy, solubilization, and interfacial tension) that have been
used to estimate asphaltene molecular weights and to probe association phenomena.
They concluded that asphaltene molecules are associated in "reversed" micelles (i.e., the
polar part of the molecule facing the center of the micelle) and asphaltenes interact
selectively with resins, although they stated the evidence on this point is subject to
alternate interpretations.

Neumann et al. (1981) using a technique called ultrafiltration, among others, determined
that asphaltenes and resins are poly-dispersed, resuspendable, spherical, and oleophilic
micelle colloids. They are oleophilic because they are stably dispersed in hydrocarbons.
They are resuspendable because, after separation by ultrafiltration, they can again be
colloidally dispersed. They are poly-dispersed because during multistage ultrafiltration
they are progressively held back on filters with different pore sizes. And they are
spherical because even solutions with relatively high concentrations (e.g., 10%) form
low-viscosity oils and not highly viscous gels, furthermore, ultramicroscopic surveys
show almost globular particles. Every colloid consists of several molecules. The
ultrafiltration of several oils showed that no filtration is achieved using filters with an
average pore diameter >350Å. Successively, smaller diameter filters separate varying
(i.e., increasing) amounts of residue. With a filter of 50Å pore diameter, all colloids can
be separated. Neumann et al. determined that it is the colloids that cause a black-brown
coloration in oils and their residual fractions.

Waxes
Some crude oils, when their thermodynamic conditions are changed (especially
5

composition and temperature), precipitate organic solids usually in the form of crystals
which are generally called waxes. Wax deposition is a serious field problem encountered
during crude oil production that causes plugging of pipelines, well tubings, and surface
and process equipment. Wax crystals change the flow behavior of crude oil from
Newtonian to non-Newtonian. The wax crystals usually lead to higher viscosity, with
increased energy consumption for pumping and a decreased pumping capacity. Wax
deposition increases the pipeline roughness which results in an increase in pressure drop.
The other effect is to reduce the effective cross sectional area of the pipe. The deposits
also cause subsurface and surface equipment plugging and malfunction, especially when
oil mixtures are transported across Arctic regions or cold oceans. Wax deposition leads
to more frequent pigging requirements. If the deposits get too thick, they reduce the
capacity of the pipeline and cause the pigs to get stuck. Wax deposition in well tubings
and process equipment may lead to more frequent shutdowns and operational problems.
Also, wax   deposition can result in severe formation damage (Sutton and Roberts,
1974). If the temperature of the fluid in the formation falls below the cloud point, wax
precipitates and may deposit in the formation pores, partially blocking or plugging the
fluid flow channels and thus restricting the flow.

The lighter components of crude oil help to keep the heavier components in solution.
These higher molecular weight solids precipitate whenever anything occurs that decreases
the carrying capacity of the fluid solvent. Crude oil is a mixture of a wide range of
hydrocarbons. The carbon number distribution of the paraffins varies from one crude oil
to another. The solubility of a specific n-paraffin in a crude oil is a strong function of its
carbon number and system temperature, as demonstrated by Nenniger and Nenniger
(1990). They showed that the solubility of C40 in a crude oil decreases by more than 2
orders of magnitude as the crude oil temperature decreases from 50 to 20 °C.
Temperature is a major driving force for the paraffin solid-liquid phase equilibria and the
subsequent separation of the two phases. The two major parameters that affect the
solubility of wax in oil are the temperature and composition of the oil. Pressure has a
lesser effect as it will be shown later. The precipitation of wax is sometimes irreversible
in that the wax, once removed from solution, is very difficult to re dissolve in the same
fluid, even after original formation temperatures are restored.

The cloud and pour points are useful in estimating the behavior of wax in oils. The cloud
point is the temperature at which wax first begins to precipitate. This causes the oil to
appear cloudy. Cloud point is a measure of the paraffinicity of a fuel oil, a high value
indicating a straight-run paraffinic oil and a low value indicating an aromatic, naphthenic,
or highly cracked oil. This point can be measured by several means in the laboratory such
as viscosity changes when temperature is decreasing, differential thermal analysis, near-
infrared spectrometry, and filtration techniques. Pour point is defined as the lowest
temperature at which the fuel will pour and is a function of the composition of the fuel.
Normally, the pour point of a fuel should be at least 10 to 15 degrees below the
anticipated minimum use temperature (Van Nostrand, 1986).

Precipitation of wax from petroleum fluids is considered to be a thermodynamic


molecular saturation phenomenon. Wax molecules are initially dissolved in a chaotic
molecular state in the fluid. At some thermodynamic state the fluid becomes saturated
6

with wax molecules, which then begin to precipitate. This thermodynamic state is called
the onset of wax precipitation or solidification. It is analogous to the usual dew point or
condensation phenomenon, except that in wax precipitation a solid is precipitating from a
liquid, whereas in condensation a liquid is precipitating from a vapor. In wax
precipitation, resin and asphaltene micelles behave like heavy molecules. When their
kinetic energy is sufficiently reduced due to cooling, they precipitate out of solution but
they are not destroyed. If kinetic energy in the form of heat is supplied to the system,
these micelles will desegregate and go back into stable suspension and Brownian motion.

Big savings can be obtained from the accurate advance prediction of wax formation.
Knowledge of the magnitude of wax deposition can lead to reduction of insulation
requirements for production and transportation systems. Also, wax problems can be
dealt with in an early stage of a project, so that sufficient thermal insulation is planned
for, instead of expensive chemical injection and loss in capacity. Process heat loads can
be reduced by increasing the efficiency of heat transfer. Capacity reduction in heat
exchangers due to blockage or vibration problems due to high velocity or flashing can be
overcome. The size of export pumps and flowlines can be reduced by having an accurate
knowledge of the effect of wax formation on crude viscosity. The minimum   pigging
frequency can be determined if the amount of wax deposition can be estimated. Also,
problems related to start-up and shutdown can be solved cost-effectively.

Conventional Hydrocarbon Phase Behavior


The phase behavior of reservoir hydrocarbons has been studied and reported extensively
for over a century. Some prominent investigators of reservoir hydrocarbon phase
behavior are Sage and Lacey, 1939, Katz et al., 1959, Standing, 1977, and McCain, 1990.
There were many other contributors to reservoir hydrocarbon phase behavior. One of the
thermodynamic tools these scientists used to describe and represent reservoir hydrocarbon
phase behavior is the familiar Pressure-Temperature Phase Diagram (P-T Diagram) or,
for reasons that will become evident later in this paper, what I will refer to from now on
as Vapor-Liquid Envelope (V-L Envelope). The V-L Envelope with the vapor quality
lines shows at a glance a lot of the phase behavior of interest of a reservoir hydrocarbon
mixture. A typical P-T Diagram of a reservoir hydrocarbon mixture is shown in Fig. 2.
The V-L Envelope consists of a bubble-point line and a dew-point line that intersect at
the critical point. The critical point is the temperature and pressure at which the
properties of the liquid and gas become identical, hence, there is only a single phase. At
conditions lying inside the envelope the fluid mixture exists as two equilibrium phases, a
vapor and a liquid. The relative amounts of the vapor and a liquid depend on
temperature, pressure, and overall composition.

The exact shape of the V-L Envelope and the critical point depends on the composition of
the fluid. Depending on the composition of the reservoir hydrocarbon mixture, naturally
occurring hydrocarbons are classified into the well known "Five Reservoir Fluids"
(McCain, 1990). The following names were assigned to these fluids which are widely
used in the Petroleum Industry (with some variations):
• Black Oils
• Volatile Oils
• Retrograde Gas
7

• Wet Gas
• Dry Gas
The type of a given reservoir fluid is important to know because many decisions
regarding oil recovery depend on it, such as, for example, fluid sampling and handling,
laboratory techniques, depletion plans, subsurface and surface equipment design and
other.

One key difference in the P-T Diagrams of the five reservoir fluids is the location of the
reservoir pressure and reservoir temperature point (Pr,Tr) relative to the critical point
(Pc,Tc). For the five reservoir fluids the location of the (Pr,Tr) point relative to the
(Pc,Tc) point is:
• Black Oils - Tr is very far to the left of Tc (Tr<<Tc)
• Volatile Oils - Tr is to the left but close to Tc (Tr<Tc)
• Retrograde Gas - Tr is to the right but close to Tc (Tr>Tc).
But, Tr<Tcrico, where Tcrico is the cricondentherm of the fluid (the highest
temperature with two coexisting phases).
• Wet Gas - Tr is far to the right of Tc (Tr>>Tc).
But, the separator conditions fall within the V-L Envelope. In other words, at
separator conditions there is production of condensate.
• Dry Gas - Tr is very far to the right of Tc (Tr>>>Tc).
The separator conditions fall outside the V-L Envelope and, as a result, at
separator conditions there is no production of condensate.
For more information regarding the five reservoir fluids, one should read McCain, 1990.

Asphaltene and Wax Phase Behavior


Asphaltene Deposition Envelope
When I started my quest to unlock the mysteries surrounding asphaltene phenomena,
about 15 years ago, one of the most fascinating questions to me was what is the
relationship between asphaltene and wax phase behavior and the conventional phase
behavior that I described briefly in the last section. For instance, from my field
experience in oil and gas production I knew that when the well-head-flowing-pressure
(WHFP) of some wells dropped below a certain point, say Po, asphaltenes started
depositing in the well and other production equipment. When the pressure was raised
above Po asphaltene deposition stopped (Leontaritis and Mansoori, 1989). This was
indicating that the onset of asphaltene flocculation and deposition was pressure
dependent. For the Prinos oil this pressure was about 90-100 atm. We also had observed
that wells that were running hotter (i.e., at higher production rates) had an onset pressure
Po lower than those wells running cooler. This indicated that the onset pressure was a
function of temperature. In other words, the formation of the asphaltene phase was
dependent on pressure and temperature.

The above behavior is what we typically call phase behavior and gives rise to the
existence of a P-T Diagram analogous to the one discussed previously for the five
reservoir fluids relative to the formation of vapor and liquid phases. This P-T Diagram of
asphaltenic fluids I have given the name Asphaltene Deposition Envelope (ADE). What
may be considered a typical ADE of a reservoir oil is shown in Fig. 3 along with the V-L
Envelope. There is clearly an overlap of the ADE with the VLE. In this P-T Diagram
8

there is no critical point shown for the fluid. The reason is there simply does not exist
one. This is a key difference between the "black oil" described earlier and an asphaltenic
reservoir fluid. To have a critical point a fluid, as defined previously, must have both a
bubble-point line and a dew-point line. Asphaltenic fluids do not have dew-point lines.
Asphaltenes simply do not vaporize at any temperature (actually, asphaltenes disintegrate
at high temperatures). Hence, the P-T Diagram of an asphaltenic reservoir fluid should
show a bubble-point line, an upper ADE boundary, and a lower ADE boundary (and, of
course, the asphaltene phase quality lines).

Some asphaltenic reservoir fluids exhibit what has been termed irreversible asphaltene
flocculation. This means that once the upper ADE boundary is crossed, e.g., during
sampling or production, some of the asphaltenes will not de-flocculate and go back into
stable suspension by simply reversing the thermodynamic path. For such reservoir fluids,
it is generally not possible to measure a lower ADE boundary because one would have to
cross the ADE to get to it. That would result in a permanent change in the composition of
the asphaltenes due to irreversible asphaltene flocculation and in the shape of the lower
(and the upper) ADE boundary. Bypassing the entire ADE to get to the lower ADE
boundary without asphaltene flocculation appears possible but it is probably of academic
interest only.

Wax Deposition Envelope


Many reservoir fluids at some frequently encountered field conditions precipitate organic
solids. For the purposes of this paper, based on the definitions given earlier, we will call
these organic solids field waxes. These waxes usually consist of a mixture of heavy
hydrocarbons such as asphaltenes, resins, paraffins, cycloparaffins, and heavy aromatics.
Wax Precipitation primarily depends on fluid temperature and composition. Pressure has
a smaller effect on wax precipitation. As with asphaltenes, the fact that waxes precipitate
at some and not at other thermodynamic states, for a given fluid, indicates that there is a
portion of the thermodynamic space that is enclosed by some boundary within which
waxes precipitate. This bounded thermodynamic space I have given the name Wax
Deposition Envelope (WDE). A typical WDE is shown in Fig. 4. The upper WDE
boundary has usually a positive slope, but it can also have a negative slope. In all cases
known to me where the WDE has been measured the upper WDE boundary was very
close to the vertical line.

The intersection of the WDE boundary with the bubble-point line is expected to be
always to the left of the onset of wax crystallization (cloud point) of the stock tank oil.
This has been the case with several reservoir fluids that I am aware of whose WDE has
been measured. This experimental fact is not expected to change because light ends,
when pressured into oil, always cause a suppression of the temperature at the onset of
wax crystallization. The actual shape of the lower WDE boundary is primarily a function
of the compositions of the intermediates and light ends of the reservoir fluid.

ADE and WDE Case Histories


The technology for measuring ADEs and WDEs only recently has been perfected.
Because the technology is new and the cost of measurements is relatively high, most
companies do not find it economical to obtain the complete ADE or WDE of their
9

reservoir fluid through laboratory measurements. Hence, preference is given to obtaining


only a few experimental data points and use them to tune phase behavior models which
calculate the remaining ADE or WDE more economically. A few cases of ADE and
WDE examples obtained in the laboratory and/or through simulation are presented next.

Asphaltene Deposition Envelopes


Prinos Oil ADE
The first Pressure-Temperature ADE (P-T ADE) ever obtained was the Prinos Oil P-T
ADE (Leontaritis and Mansoori, 1989). The actual laboratory data, which were obtained
by the Institute of Petroleum of France (IFP) and were supplemented by field
observations, were collated together to form the ADE shown in Fig. 5. It is not known
whether the actual Prinos ADE had closure at high or low temperatures or both. An
improved version of the Thermodynamic-Colloidal Model (T-C Model) by Leontaritis,
1993, predicts closure of the Prinos ADE at high temperatures. Also, the T-C Model
predicts the low ADE boundary for the Prinos oil at much lower pressures than those
shown in Fig. 5. The fact that the measured lower ADE boundary shown in Fig. 5 is so
close to the bubble point line could be explained by the loss of some of the most polar
and heavy molecules due to sample handling. Sample handling techniques at that time
were not as advanced as they are today. Furthermore, asphaltene flocculation in the
Prinos oil had been determined to be irreversible because we could not reproduce the
reservoir fluid asphaltene phase behavior by using reconstituted live oil from separator
liquid and gas.

South American Oil - ADE


Fig. 6 shows the P-T ADE of a South American reservoir fluid obtained by modern
sample handling and laboratory techniques. Two points were obtained on the upper and
two on the lower ADE boundaries. Also, the amount of asphaltene phase formed versus
pressure, at reservoir temperature, was also measured. Additional measurements, made
after the ones in Fig. 6 and shown in Fig. 7, on a fluid obtained from another payzone of
the same reservoir, point out clearly that closure of the ADE occurs at high temperatures.
In Fig. 7, at high temperatures, the upper ADE has a negative slope whereas the bubble-
point line shown in Fig. 6 has a positive slope. The actual intersection point between the
upper ADE boundary and the bubble-point line is not known, but it was estimated to be
around 370 °F.

South American Oil - Simulated ADE


Fig. 8 shows the simulated P-T ADE of a North American reservoir fluid. AsphWax Oil
Company is a fictitious name. For this reservoir fluid the upper and lower onset of
asphaltene flocculation pressure at reservoir temperature was obtained using modern
laboratory technology. Also, state of the art sampling technology was used. Two
asphaltene sedimentation measurements were made inside the ADE at reservoir
temperature, one above and the other below the bubble-point pressure. An improved
version of the Thermodynamic-Colloidal Model (Leontaritis, 1993) was used to simulate
the entire ADE and VLE. It should be noted that a standard black oil study was available
for the subject reservoir fluid which was used to tune the VLE portion of the model.
Although no laboratory data are available to substantiate this, the T-C Model calculates a
lower ADE boundary that is "bulging out" toward higher temperatures. This is the first
10

time that I observed this behavior and I am taking steps to reproduce it experimentally.

To highlight some of the salient features of this thermodynamic diagram, I have plotted
additional simulation results which are shown in Fig. 9 and 10. Fig. 9 shows the effect of
temperature on asphaltene phase volume formation as the ADE is traversed horizontally
at 200 atm. This is the volume of asphaltene phase per mole reservoir fluid that would
form in a PVT cell at thermodynamic and sedimentation equilibrium. The upper liquid
phase in the PVT cell also contains asphaltene particles (usually a bigger amount than the
lower asphaltene phase) that do not settle at system conditions but continue to move
around in a random Brownian motion. Fig. 10 shows the effect of pressure on asphaltene
phase volume formation as the ADE is traversed vertically at 340 °K.

Wax Deposition Envelopes


From the early twentieth century scientists and engineers knew that the addition of light
ends to crude oil suppresses the tendency of waxes to precipitate (Reistle, 1932).
Recently considerable attention has been given on the effect of light ends on wax phase
behavior (Won, 1986, Weingarten, et al., 1988). Fig. 11 demonstrates the dramatic
impact light ends have on the onset of wax crystallization. The onset of wax
crystallization of this fluid was supressed by about 40 °F by pressuring separator gas into
the stock tank oil. Pressure has a smaller but definite effect on the onset of wax
crystallization. Fig. 12 shows about a 10 °F rise in the cloud point for a 3000 psi pressure
increase of a synthetic wax-kerosene mixture. Because of the regularity of seeing a
similar pressure effect on the onset of wax crystallization of several synthetic and natural
hydrocarbon mixtures, it was very tempting to conclude that pressure always causes a rise
in the onset of wax crystallization. Until, I bumped into a live fluid for which pressure
caused a decline on the onset of wax crystallization. Fig. 13 shows the effect of pressure
on the onset of wax crystallization for three live oils. Pressure suppresses the onset of
wax crystallization of oil B but raises the onset of wax crystallization for Oils A and C.

North American Recombined Live Oil - WDE


Fig. 14 shows the WDE of a recombined reservoir fluid obtained with state of the art
laboratory techniques. The fact that the fluid is recombined casts some doubt as to how
representative the measured WDE is of the real reservoir fluid. Any changes in the
composition of the heavy fractions of the fluid, no matter how small, are expected to
impact the shape of the WDE.

North Sea Live Oil - Simulated WDE


Fig. 15 shows the simulated P-T WDE of a North Sea live oil. AsphWax Oil Company
remains a fictitious oil company name. The WaxModel utilized in this simulation is an
improved version of the one published by Loganathan et al., 1993. A standard black oil
PVT study was available for this fluid. The PVT study was used to tune the VLE
predictions of the WaxModel. The only wax data available for tuning the model were the
cloud point of the STO at 1 atm and the onset of wax crystallization temperature of the
reservoir fluid at 200 atm. By using these two wax measurements, the asphaltene/resin
and wax contents of the fluid, and the black oil PVT study the WaxModel was able to
calculate the complete WDE shown in Fig. 15.
11

Fig. 16,17,18, and 19 illustrate the predictions of the WaxModel inside the WDE (both
above and below the bubble-point line). Figure 16 shows the amount of wax formed at
200 atm as the temperature drops. Figures 17 and 18 show the same thing but at 50 atm
and 1 atm pressure respectively. Figure 19 shows the amount of wax formed at 280 °K as
the pressure drops from 200 atm to 1 atm. The above-mentioned calculations are
different traverses of the WDE shown in Figure 15.

Hypothesis - Unification of ADE and WDE


Although most (in not all) scientists and engineers would agree today that asphaltene and
wax phase behavior differ from each other, based on current know-how, there is I believe
a good reason for one to think that there is some relationship between the two. Both
cases concern organic solids depositing from a hydrocarbon mixture. There are, of
course, compositional changes which account for the variation in phase behavior, as
presented in this paper. However, it is reasonable for one to expect to find a continuum
of phase behavior from asphaltenic to waxy fluids because nature rarely demonstrates
discontinuities of this kind. I am hypothesizing that asphaltene and wax phase behavior
represent two extremes (left and right) of solid-liquid phase behavior that can be
attributed to extreme compositional differences of the naturally occurring hydrocarbons.
One then should expect to find reservoir fluids with compositions that display
intermediate type of solid-liquid phase behavior. Let me elaborate further on how this
intermediate phase behavior might manifest itself.

As demonstrated in this paper, the ADE and WDE summarize asphaltene and wax phase
behavior. As a result, they are convenient tools for probing into and demonstrating the
similarities and differences of asphaltene and wax phase behavior. Experience shows that
asphaltene flocculation occurs primarily at high temperatures, whereas wax crystallization
occurs primarily at low temperatures. This means that the kinetic energy of the heavy
molecules and micelles plays only a small role in asphaltene flocculation but a major one
in wax crystallization. In wax crystallization the large molecules and micelles, as shown
in Fig. 1, simply aggregate together and form a solid phase because their kinetic energy is
not enough to overcome the attractive forces that exist when they bump into each other.
When the temperature of the system is raised the kinetic energy of the molecules and
micelles is enough to keep them apart. Light ends (light hydrocarbons) have an
analogous effect on the behavior of molecules and micelles. They dilute the large
molecules and micelles thus allowing them to stay apart at lower temperatures.

On the other hand, in asphaltene flocculation the temperature of the system and their
kinetic energy is so high that aggregation of the molecules and/or micelles is not possible.
When they bump onto each other the attractive force field that surrounds them (these are
the traditional van der Waals-London dispersion forces) is not enough to keep them
together. In asphaltene flocculation the nature of the liquid that acts as the "capacitance"
that keeps the micelles apart is very crucial to their stability. As the thermodynamic
conditions change the nature of the liquid (and its capacitance) changes and as a result the
force equilibrium that keeps the micelles apart changes. Experiments have shown that
resins play a key role in the stability of asphaltenes (Swanson, 1940, Murzakov, 1980,
Leontaritis et al., 1994). It is behooving then to conclude that tracking the compositional
changes of the resins as the thermodynamic conditions of the system vary should help in
12

predicting the stability of asphaltene micelles. The T-C Model is based on this concept
(Leontaritis and Mansoori, 1987).

If we accept that natural hydrocarbon systems that occur in the earth's crust exhibit phase
behavior that is a continuum between the two extremes of asphaltene and wax phase
behavior, we should then expect to find some real hydrocarbon mixtures whose phase
behavior is in-between that of asphaltenes and waxes and, as a result, indistinguishable.
An interesting question is, how would this continuum of phase behavior manifest itself?
To scope out a possible answer to the above question, we should direct our attention to
the new thermodynamic tools presented in this paper, the ADE and WDE.

In general, we expect the ADE to face to the left (i.e., have an opening toward low
temperatures) and the WDE to face to the right (i.e., have an opening toward high
temperatures). It is behooving to assume then that the deposition envelopes between the
two extremes will have intermediate shapes. This is illustrated in Fig. 20. At high
temperatures the deposition envelope faces to the left (like the ADE). Moving to lower
temperatures the deposition envelope smoothly changes shape until it becomes similar to
WDE which faces or curves to the right. Experimental evidence to date shows that solids
form always to the left of the deposition envelopes (although there is some invalidated
evidence to the contrary). My experimental and modeling experience indicates that solids
should occur always to the left of the ADE and WDE regardless of the shape. I suppose
many reliable future laboratory tests with various fluids obtained from around the world
are needed before one can be absolutely sure of this.

Conclusions
Asphaltene and wax phenomena differ appreciably. Asphaltenes flocculate (form flocs)
whereas waxes crystallize or solidify (precipitate due to saturation). The stability of
asphaltenes depends mainly on the nature of their suspension medium, the liquid. The
stability of waxes depends mainly on their temperature (kinetic energy) and concentration
(saturation effect). These differences give rise to significant variations in asphaltene and
wax phase behavior. Two new thermodynamic diagrams were introduced to summarize
and represent the asphaltene and wax phase behavior of reservoir hydrocarbons. The
ADE and WDE complement the existing thermodynamic diagrams. It was noted that
reservoir fluids that contain significant amount of asphaltenes do not exhibit a dew-point
line because asphaltenes do not vaporize at any temperature or pressure. As a result,
asphaltenic fluids do not have critical points.

Furthermore, it was suggested that asphaltene and wax phase behavior are two extremes
of solid-liquid phase behavior. Whether the fluid displays asphaltene or wax phase
behavior depends on the composition of its heavy fractions. It was hypothesized that
since a continuum of reservoir fluid compositions is expected to occur in the earth's crust
ranging from very asphaltenic to very waxy fluids, the thermodynamic diagrams of these
fluids should be in-between the two extremes of ADE and WDE. More deposition
envelope measurements with a variety of reservoir fluids will shed some light on this
point in the not distant future.
13

References
Katz, D. L., Cornell, D., Kobayashi, R., Poettmann, F. H., Vary, J. A., Elenbaas, J. R.,
Weinaug, C. F., "Handbook of Natural Gas Engineering," McGraw-Hill Book
Company, New York (1959).

Koots, J. A. and Speight, J. C., “Relation of Petroleum Resins to Asphaltenes,” Fuel, 54,
p. 179-184 (1975).

Leontaritis, K.J., "Application of a Thermodynamic-Colloidal Model of Asphaltene


Flocculation," The Symposium of Solids Deposition, Area 16C of Fuels and
Petrochemical Division, AIChE Spring National Meeting and Petroleum
Exposition, Houston, Texas, March 28-April 1, 1993.

Leontaritis, K. J. and Mansoori, G. A., "Asphaltene Deposition. A Survey of Field


Experiences and Research Approaches," International Journal of Petroleum
Science and Engineering, 1, pp. 229-239, 1989.

Leontaritis, K.J. and Mansoori, G. A., "Asphaltene Deposition During Oil Production and
Processing: A Thermodynamic Colloidal Model," SPE Paper No. 16258, SPE
International Symposium on Oilfield Chemistry, February 4-6, 1987, San
Antonio, TX.

Leontaritis, K.J., Amaefule, J.O. and Charles, R.E., "A Systematic Approach for the
Prevention and Treatment of Formation Damage Caused by Asphaltene
Deposition," SPE Paper No. 23810, Formation Damage Symposium, February 26-
27, 1992, Lafayette, Louisiana.

McCain, W. D. Jr., "The Properties of Petroleum Fluids," PennWell Publishing


Company, Tulsa, Oklahoma (1990).

Murzakov, R. M., Sabanenkov, A. A., and Synmyaev, Z. I., “Influence of Petroleum


Resins on Colloidal Stability of Asphaltene-Containing Disperse Systems,”
Translation from Khimiya i Tekhnobgiya Topliv i Mase, No. 10, pp. 40-41,
(October, 1980).

Narayanan, L., Leontaritis, K.J., and Darby, R., "A Thermodynamic Model for Predicting
Wax Precipitation from Crude Oils," The Symposium of Solids Deposition, Area
16C of Fuels and Petrochemical Division, AIChE Spring National Meeting and
Petroleum Exposition, Houston, Texas, March 28-April 1, 1993.

Nenniger, J. and G. Nenniger, “Optimizing Hot Oiling/Watering Jobs to Minimize


Formation Damage,” SPE Paper # CIM/SPE 90-57, presented at the International
Technical Meeting in Calgary, June 10-13, (1990).

Neumann, H. J., Paczynska-Lahme, B., and Severin, D., “Composition and Properties of
Petroleum,” Ferdinand Enke Publishers, Stuttgart, German, (1981).
14

Pfeiffer, J. P. and Saal, R. N.J., Asphaltic Bitumen as a Colloid System, J. Phys. Chem
44, 139. (1940).

Reistle, C. E. Jr., "Paraffin and Congealing-Oil Problems," Bulletin 348, USBM (1932).

Sage, B. H. and Lacey, W. N., "Volumetric and Phase Behavior of Hydrocarbons," Gulf
Publishing Company, Houston, Texas (1939).

Speight, J. G., Wernick, D. L., Gould, K. A., Overfield, R. E., Rao, B. M. L., and Savage,
D. W., “Molecular Weight and Association of Asphaltenes: A Critical Review,”
Revue De L’Institut Francouis Du Petrole, Vol 10, No. 1, (1985).

Standing, M. B., "Volumetric and Phase Behavior of Oil Field Hydrocarbon Systems,"
Society of Petroleum Engineers of AIME, Houston, Texas (1977).

Sutton, G. D., and L. D. Roberts, “Paraffin Precipitation During Fracture Stimulation,” J.


Pet. Technol., Sep., 997 (1974).

Swanson, J., "A Contribution to the Physical Chemistry of the Asphalts,"J. Phys. Chem.,
46, p. 141, (1942).

Van Nostrand Scientific Encyclopedia, 5th ed., Van Nostrand Reinhold Company,
NewYork (1986).

Weingarten, J. S., and Euchner, J. A., “Methods for predicting Wax Precipitation and
Deposition,” SPE 15654, 61st Annual technical Conference & Exhibition, New
Orleans, Oct. 5-8, (1986).

Won, K.W., “Continuous Thermodynamics for Solid-Liquid Equilibria : Wax Formation


from Heavy Hydrocarbon Mixtures,” Paper 27A presented at AIChE Spring
National Meeting. Houston, TX. March 26, (1985).
15
16

LIQUID PHASE

ASPHALTENE PHASE

Figure 1. Physical Model of Asphaltenic Oils


17
18

Cricondenbar

CP

Liquid
Vol%
Liquid

100%
Pressure

ne
t Li 75%
in
-Po
le Vapor
Cricondentherm
bb 50%
Bu
ne
Li
int

25%
o
-P
w
De

0%

Temperature

Figure 2. Typical Reservoir Hydrocarbon P-T Phase Diagram


19
20

(P res, Tres)
Liquid Phase •

ADE Up
p er Bound
ary
Bubble-Point Line

Liquid + Asphaltene Phases


Pressure

Liquid + Vapor + Asphaltene Phases

Liquid + Vapor Phases


ry
u nda
Bo
er
ow
EL
AD

Temperature

Figure 3. Typical P-T Asphaltene Deposition Envelope


21
22

Bubble-Point Line

Liquid Phase
Solid+ Liquid Phases
Pressure

Liquid + Vapor Phases

Solid+Liquid+Vapor Phases

Temperature

Figure 4. Wax Deposition Envelope


23

Figure 5

P-T Asphaltene Deposition Envelope


Prinos Reservoir Fluid
(Data from Well PA-7, Zone A1)
Pressure, Atm

200

ry
nda
r Bou Bubble Point Line
ppe
EU ASPHALTENE
AD
100 DEPOSITION
ENVELOPE

ar y
ound
er B
Low
ADE

0 100 200

Temperature, °C
24

Figure 6. Asphaltene Deposition Envelope


South American Reservoir Fluid

7000

Upper ADE Boundary

6000 1.0*
2.0
3.0
Pressure, psig

5000
4.0
e
tion Lin 3.0
Satura
4000

3000
ry
o unda
DE B
er A
Low
2000
140 180 220 260 300

Temperature, °F

* Mls of asphaltene phase per mole of reservoir fluid.


25

Figure 7. Effect of Temperature on


Onset of Asphaltene Flocculation
South American Reservoir Fluid

9000

8500

8000

7500

7000

6500

6000
140 160 180 200 220 240 260

Te mpe rature , °F
26

Figure 8. Asphaltene Deposition Envelope


AsphWax Oil Company Live Oil
Reservoir Pressure, 350.0 atm
450 Reservoir Temperature, 344.27 °K
No Solids
400

350

300 Solids
250

200 No Solids

150 Solids
100
50

0
280 330 380

Te mpe rature , °K
LowerOnsetP Bubble P UpperOnse tP
27

Figure 9. Asphalte ne Phase Volume vs. Te mpe rature


AsphWax Oil Company Live Oil at 200 atm

420

Reservoir Pressure, 350.0 atm


400 Reservoir Temperature, 344.27 °K

380

360

340

320

300
0 0.5 1 1.5 2 2.5 3 3.5
Asphalte ne Phase Vo lume , cc
28

Figure 10. Asphalte ne Phase Volume vs. Pre ssure


AsphWax Oil Company Live Oil at 340 °K

450
Reservoir Pressure, 350.0 atm
400 Reservoir Temperature, 344.27 °K
Bubble Point Pressure, 279.16 atm at 340 °K
350

300

250

200

150

100

50

0
0 2 4 6 8
Asphalte ne Phase Vo lume , cc
29

Figure-11 Effect of Light Ends on Wax Phase Behavior

130
Onset of wax crystallization temperature,

120

110

100
degrees Fahrenheit

90

80

70

60
0 1000 2000 3000 4000 5000

Bubble point pressure, psig


(at 210 degrees Fehrenheit)
30

Figure 12. Pre ssure -Te mpe rature Effe cts on


Onse t of Wax Crystallization
Ke rose ne -Candle Wax Synthe tic Mixture

6000

5000

4000

3000

2000

1000

0
68 70 72 74 76 78 80 82 84
Onse t T e mpe rature , °F
31

Figure 13. Upper WDE Boundaries


of Three Reservoir Fluids

140

130

120

110

100

90

80

70
2000 3000 4000 5000 6000 7000 8000

Pre ssure , psig


Oil A Oil B Oil C
32

Figure 14. Wax Deposition Envelope


North American Recombined Reservoir Fluid

3500

3000

2500
Pressure, psig

2000

1500

1000

500

0
30 40 50 60 70 80 90 100 110

Temperature, °F
Onset Pressure, psig BP Pressure, psig
33

Figure 15. Wax Deposition Envelope


AsphWax Oil Company Live Oil
Reservoir Pre ssure, 280.0 atm
300 Reservoir Te mperature, 338.0 °K

250
No Solids
Solids
200

150

100 No Solids

50 Solids

0
250 270 290 310 330 350

Te mpe rature , °K
Onset Pressure Bubble Point Pressure
34

Figure 16. Wax Wt. Fraction of Liquid


vs. Te mpe rature
AsphWax Oil Company Live Oil at 200 atm

0.100 Reservoir Pressure, 280.0 atm


Reservoir Temperature, 338.0 °K
0.080

0.060

0.040

0.020

0.000
240 250 260 270 280 290 300
Te mpe rature , °K
35

Figure 17. Wax Wt. Fraction of Liquid


vs. Te mpe rature
AsphWax Oil Company Live Oil at 50 atm

Reservoir Pressure, 280.0 atm


0.15 Reservoir Temperature, 338.0 °K

0.12

0.09

0.06

0.03

0
250 270 290 310 330 350
Te mpe rature , °K
36

Figure r 18. Wax Wt. Fraction of Liquid


vs. Te mpe rature
AsphWax Oil Company Stock Tank Oil at 1 atm

0.3 Reservoir Pressure, 280.0 atm


Reservoir Te mperature, 338.0 °K

0.25

0.2

0.15

0.1

0.05

0
250 270 290 310 330 350

Te mpe rature , °K
37

Figure 19. Wax Wt. Fraction of Liquid


vs. Pre ssure
AsphWax Oil Company Live Oil at 280 °K
Reservoir Pressure, 280.0 atm
0.200 Reservoir Temperature, 338.0 °K

0.160

0.120

0.080

0.040

0.000
0 50 100 150 200 250
Pre ssure , atm
View publication stats

38

Figure 20. Unification of WDE and ADE

WDE Behavior ADE Behavior


Pressure

Low T High T

Anda mungkin juga menyukai