Está en la página 1de 12

PETROLEUM SOCIETY

CANADIAN INSTITUTE OF MINING, METALLURGY & PETROLEUM

PAPER 2007-144

Heavy Oil Waterflooding: Effects of Flow Rate and Oil Viscosity


A. MAI
University of Calgary, TIPM Laboratory

A. KANTZAS
University of Calgary, TIPM Laboratory
This paper is to be presented at the Petroleum Societys 8th Canadian International Petroleum Conference (58th Annual Technical Meeting), Calgary, Alberta, Canada, June 12 14, 2007. Discussion of this paper is invited and may be presented at the meeting if filed in writing with the technical program chairman prior to the conclusion of the meeting. This paper and any discussion filed will be considered for publication in Petroleum Society journals. Publication rights are reserved. This is a pre-print and subject to correction.

Abstract
Many countries in the world contain significant heavy oil deposits. In reservoirs with viscosity over several hundred mPas, waterflooding is not expected to be successful due to the extremely high oil viscosity. In many smaller, thinner reservoirs or reservoirs at the conclusion of cold production, however, thermal enhanced oil recovery methods will not be economic. Waterfloods are relatively inexpensive and easy to control; therefore they will still often be employed even in high viscosity heavy oil fields. This paper presents experimental findings of waterflooding in laboratory sand packs for two high viscosity heavy oils: 4650 mPas and 11500 mPas, at varying water injection rates. The results of this work show that capillary forces, which are often neglected due to the high oil viscosity, are in fact important even in heavy oil systems. At low injection rates, water imbibition can be used to stabilize the waterflood and improve oil recovery. Waterflooding can therefore be a viable non-thermal enhanced oil recovery technology even in fields with very high oil viscosity.

Introduction
The Canadian deposits of heavy oil and bitumen are some of the largest in the world. Our conventional oil reserves are now steadily declining, while the global energy demand continues to increase, along with a higher uncertainty about foreign oil sources. As a result, the Canadian oil sands will help Canada to remain an important energy source for the world in future generations. Heavy oil is a special class of this unconventional oil that has viscosity ranging from 50 50000+ mPas. Heavy oil reservoirs are often found in high porosity, high permeability, unconsolidated sand deposits. At reservoir conditions, the oil may contain dissolved solution gas, thus some oil can be initially recovered using the energy from heavy oil solution gas drive. At the end of primary production, a significant fraction of oil still exists for potential secondary recovery. Many of these reservoirs are small and thin or segmented, making them poor candidates for expensive thermal enhanced oil recovery strategies.

Waterflooding is often employed at least initially in these heavy oil reservoirs after primary recovery is finished. Water injection can be used to re-pressurize the reservoir and displace oil to producing wells. In these applications, it is very important to understand the forces that are present in the reservoir, and how they can be used to properly design a heavy oil waterflood. This work presents the results for water injection into laboratory sand packs containing gas-free heavy oil of varying viscosity. The responses for different waterfloods are compared in order to understand the mechanisms by which oil can be recovered by water injection.

Isr =

(M 1) w D 2 C*k wor

...................................................................... (1)

The mobility ratio (M) is defined as:


M= k wor o ................................................................................... (2) k oiw w = the injection velocity w = viscosity of water o = oil viscosity D = diameter of the core = interfacial tension kwor = permeability to water at the irreducible oil saturation Sor koiw = permeability to oil at the connate water saturation Swi and C* = wettability constant.

Where,

Theory
Waterflooding of oil reservoirs is a well-recognized technique for oil recovery after primary production. In conventional oil, waterflooding theory has been well documented1. The inherent assumption in conventional oil waterflooding theory is a similarity in viscosity between oil and water2,3. In heavy oil applications this is not the case, thus even concepts like oil/water relative permeability do not have the same meaning in heavy oil reservoirs. However, practitioners often still attempt to apply the same theoretical understanding to their fields. There has been some limited experience documented for waterfloods in heavy oil reservoirs4-7, but in general the mechanism of viscous oil recovery by waterflooding has not been explored. Waterflood recoveries are known to be low for high viscosity heavy oil, due to the adverse mobility ratio between oil and injected water. Despite the presumed inefficiency of this process, in many heavy oil fields waterflooding is still commonly applied since it is relatively inexpensive and field operators have years of experience designing and controlling waterfloods. At the end of a conventional oil waterflood, residual oil is left in place due to reservoir heterogeneities or capillary trapping. In laboratory core flood experiments, capillary bypassing is the main mechanism responsible for trapping of oil2,8. This is not the case in heavy oil. In heavy oil reservoirs, the high oil viscosity (and hence the poor mobility ratio between displacing and displaced fluids) is the main cause for oil bypassing and residual oil at the end of the waterflood. Previous investigations have therefore focused on the oil/water mobility ratio, and how it relates to viscous fingering or instability of the displacing water front9,10.

C* has different values for varying rock wettability, which indicates that the effect of imbibition on the growth of viscous fingers is different in oil wet vs. water wet porous media. In smaller diameter cores, there is less potential for fingers to grow so in the field, the effect of instability may be more pronounced than in a linear core system. Isr is also directly proportional to the fluid mobility; in heavy oil systems the mobility ratio is very large, which leads to very high values of Isr (i.e. very unstable floods). At the onset of instability, Isr was found to be9 2, or 13.56. When Isr < 13.56, the displacement is stable, indicating that viscous fingers will not grow. When Isr > 13.56 the displacement is deemed fully unstable. In the transition zone (13.56 < Isr < 1000) the flood is becoming increasingly unstable, and breakthrough recovery decreases rapidly as Isr increases. In the pseudo stable region of Isr > 1000 the recovery tends to become constant again9. In this range, displacement is actually so unstable that a single finger dominates flow; most of the injected fluid is passing through this finger, and recovery is low and relatively independent of injection rate. Instability theory dictates that before Isr = 1000, the displacement rate determines the finger properties. At a high injection rate in an unstable system, the finger wavelength will be short, thus numerous fingers will form, however at a low rate the finger wavelength will be long, thus only a few fingers can form in the porous medium9-12. Multiple fingers lead to a higher degree of instability. Therefore it is recommended to perform experiments more slowly under unstable conditions, in order to limit the generation and growth of fingers. Peters and Flock9 stressed the importance of the wettability number. This number gives an indication of the ability of the porous medium to imbibe the displacing water, which can help to stabilize the flood front. They found that in water-wet media, the imbibition forces are strong, thus the wettability number will be large (C* = 306.25). For oil-wet media, the wettability number was found to be much lower (5.45). This is because under drainage, water will only move through the largest channels, so the front cannot be stabilized by additional flow into smaller pores. Bentsen10 has also derived a different version of the instability number based upon force potentials rather than velocity potentials. His version of instability number is
2

Instability
In heavy oil waterflooding, water is displacing more viscous oil, thus the displacement front may become unstable. When this happens, viscous fingers are said to have formed. This will cause premature breakthrough of the displacing phase, and reduce the breakthrough oil recovery. Peters and Flock9 identified the parameters controlling the stability of the system as: mobility ratio, displacement velocity, system geometry and dimensions, capillary and gravitational forces, and system permeability and wettability. Their work was focused on performing stability analysis in order to identify the conditions under which a frontal perturbation will grow to become a viscous finger. The instability number (Isr) for a horizontal 1-D system, as defined by Peters and Flock9, is as follows:

proportional to the one proposed by Peters and Flock9, with an additional factor to take into consideration the larger size difference of water and oil fingers. Bentsen et al.12 also developed the theory further to predict the recovery at breakthrough for stable displacement and pseudo-stable (Isr > 1000) displacement only. The theory of instability is basically a balance of forces. In a displacement, if the combined forces of gravity and capillarity are greater than the viscous force, then the displacement will be stable. If the reverse is true then the displacement will be unstable, and the degree of instability depends on the rate of injection, with all else being equal. In heavy oil systems, the difference between oil and water viscosity is so great that Isr will always tend to be very large. This theory indicates a dominance of viscous forces during waterflooding, and explains the low recovery expected.

other pores, and displacement of viscous oil will tend to occur much more slowly. Several researchers19-21 have observed that recovery for fixed volumes of water increases in a manner that is proportional to t . In capillary-driven processes, recovery is due to imbibition. Thus, the imbibition rate can be shown to be:
dR 1 ......................................................................................... (3) dt t This implies that the imbibition rate or oil production rate should be high at first, and then should decrease with time. At early times during injection into heavy oil, however, water is displacing high viscosity heavy oil, so the viscous forces are expected to be dominant over capillary forces. In summary, imbibition rate decreases with time, increasing permeability and increasing oil viscosity. Upon consideration of conventional knowledge regarding imbibition, it is not expected that this parameter will hold great importance in high permeability oil sands. This has led to the common assumption that capillary forces and imbibition are insignificant in heavy oil systems, in relation to the effect of the viscous forces. The validation of this assumption is one of the main focuses of this research.

Imbibition
In the presence of multiple immiscible fluids in porous media, the distribution of the fluids at equilibrium is governed by capillary foces1,2. Specifically, during water injection into a water wet porous medium (imbibition), capillary forces compete with viscous forces for determining the pathways through which water will travel. Imbibition is therefore an important phenomenon during water injection. There are many factors controlling imbibition, such as: wettability, permeability, pore size, viscosities of the imbibing and displaced fluids, and the initial water and oil saturations of the rock13. Wettability is by far the most important parameter in imbibition, as evidenced by the contact angle in the YoungLaplace equation1. Wettability controls which fluid will be spontaneously imbibed into the porous medium. The strength of wettability will also influence the rate of imbibition. Even in viscous heavy oil reservoirs, the porous medium is normally water wet14. Li and Horne15 have shown that the capillary pressure is expected to decrease for rocks with higher permeability, since permeability is related to the average pore size in the rock. In separate studies they also observed that the imbibition rate is higher for lower permeability rocks16. Rock permeability is therefore another important parameter for quantifying the effect of imbibition. In unconsolidated oil sands, permeability tends to be higher than in conventional oil reservoirs, therefore the rate of imbibition will be lower in this case. Zhou et al.17 found that both the imbibition rate and the final recovery by imbibition decreases with initial water saturation. This is based on the theory of capillary pressure; at increasing water saturation, the water is expected to exist in the smallest pores and some of the larger pores, therefore capillary pressure is lower and imbibition will occur more slowly. In heavy oil systems, the capillary pressure is less well defined. At the point of water breakthrough, water has traveled through the leastresistant pore pathways, creating a channel of high water saturation. In other portions of the core, however, the oil was by-passed and therefore the condition of oil and irreducible water still exists. Fischer et al.18 have also shown that the imbibition rate is a function of oil viscosity. As oil viscosity increases, imbibition rate decreases. This result is important especially in heavy oil systems, where the oil and water do not have similar mobility. In order for imbibition to occur, oil has to be displaced into

Experimental Procedure
The sand used in all experiments was Lane Mountain 70 sand. The sand was wet packed with methanol. After the sand was packed, it was CT (Computerized Tomography) scanned to determine that the pack was uniform. The sandpack was then drained and dried overnight with compressed air, and nitrogen was used to remove any residual methanol. The sandpack was then scanned again to ensure that it was dry. Once it was established that it was dried, gas expansion was performed on the sandpack to determine its porosity. The sandpack was then left under vacuum for approximately a day, after which brine saturation and brine permeability were determined. The sandpack was then flooded with oil until irreducible water saturation is reached. During this stage, the pressure at the inlet and outlet of the sandpack was monitored. The sandpack was left undisturbed for about a day to allow equilibrium to be reached. Waterfloods were then carried out at the specified flow rates. Again, the inlet and outlet pressures were recorded periodically. The products of the waterflood was also collected and analyzed for oil and water content with low-field NMR and separation with toluene. All floods were performed at ambient temperature (23C) and with applied overburden on the sand.

Results
In a gas-free conventional oil sandpack, it is commonly expected that as water is injected, oil is continuously produced until breakthrough. After this point very little extra oil is recovered and virtually all the water injected is produced. As mentioned earlier, the oil trapped in the porous medium at the end of the waterflood is due to capillary forces. It is for this reason that in strongly water wet porous media the recoveries from waterflooding and from imbibition are essentially the same2.

The behavior of heavy oil waterfloods is distinctively different. Due to the fact that heavy oil is considerably more viscous than the water, injection of a less viscous fluid with high mobility to recover heavy oil with limited mobility leads to viscous fingering. This results in early water breakthrough and reduces the efficiency of the waterflood. The recovery profile for a heavy oil waterflood is shown in Figure 1. It is obvious that the waterflooding response is very different for conventional and heavy oil. For heavy oil breakthrough of injected water occurs early, as evidenced by the rapidly rising water cut at early times in Figure 2. After breakthrough oil is still being produced along with high water-cut. At first glance, this response is similar to waterflooding of conventional oil in an oil wet porous medium22,23. However, the sand used in these experiments is clean quartz, and during the process of the experiments, the sand is always first exposed to brine. Thus, these elements should ensure that the sand is water wet9. Additionally, the pressure profile does not match that of a waterflood in an oil-wet core (i.e. a drainage process). At early times pressure builds up upon constant rate water injection, and after breakthrough pressure decreases down to very low values24. If the recovery profile in Figure 1 was simply a reflection of rock wettability, then the injection pressure would have to remain high in order to access progressively smaller pores. It can therefore then be concluded that this characteristic is due to some different mechanism rather than the porous medium being oil wet. After breakthrough water has found continuous pathways from inlet to outlet, thus further injection should continue to follow these paths of least resistance. Along this path, water extracts oil and thus improves recovery. As a result oil is produced with extremely high water cut, as shown in Figure 2. It is extremely important to determine the mechanisms contributing to oil production after water has broken through. The possible forces acting on the porous medium fluids are viscous, capillary, gravitational and inertial. For the experiments performed in this study, the floods were performed horizontally and the diameter of the sandpack is too small (3.81 cm) to render any gravitational effects, thus they can be neglected. At the low velocity flow rates utilized in this work, inertial (shear) forces are also negligible. The two competing mechanisms are then viscous and capillary forces. Various experiments were performed to extrapolate the relative important of viscous and capillary forces.

density of 0.9815 g/cm3 and a viscosity of 11500 mPa.s. The brine used in all these experiments has a density of 1.0487 g/cm3. These properties were measured at 23C. The analysis of waterflooding results for conventional oils often involved the prediction of breakthrough recoveries as a function of capillary number1,2, in which recovery increases as capillary number increases. Capillary number, Nca, is defined as:

N ca =

w .................................................................................. (4) cos

For heavy oil waterfloods it is not simple to identify the occurrence of breakthrough. Generally, breakthrough is identified as the point when water was first produced. However, in a heavy oil waterflood, water is produced very early on in the experiment, at which point the recovery is still very small. Thus a new definition of breakthrough must be established in order to provide reasonable production. From these experiments it was seen that as water was injected, heavy oil with limited mobility could not be produced as fast as the injected rate, thus the pressure in the porous medium consistently built up. At a certain point, water will finger through and the pressure will decline. At the point where pressure begins to decline, this is identified as the break through point. The breakthrough recoveries along with their capillary numbers, as defined by equation (4) are shown in the Table 2. These results show that as injection rate increases, recovery at breakthrough decreases, except for the cases of 1 mL/hr. They also show that recovery at breakthrough is better for the less viscous oil. At this range of capillary numbers, the recoveries in conventional oil are significantly higher than the values in Table 21,2. This demonstrates the main difference between conventional oil and heavy oil waterfloods: at water breakthrough residual heavy oil was bypassed due to adverse mobility ratio rather than from capillary trapping. Common knowledge of waterflooding of conventional oil often dictates that to improve oil recovery, higher injection rate should be employed. Due to similar viscosities between the displacing and the displaced phase, the injection front advances almost as a piston. This means that in general a large fraction of the porous medium was swept with the residual oil exists in a microscopic form as isolated ganglia. In a strongly water wet medium, water will invade smaller pores preferentially due to imbibition. If injection rates are sufficiently high, water can enter both large and small pores, leading to a more efficient displacement. This ensures that the remaining oil in the porous medium is lower than the case of smaller injection rate, thus improving recovery. However, in heavy oil waterfloods, instability at the flood front would develop into viscous fingers with higher injection rates leading to less efficient displacement9-12. Macroscopically, a significant area of the porous medium was unswept due to severe bypassing. Thus, a large portion of the remaining oil is still continuous. In order to analyze the results of heavy oil waterfloods, viscous fingering must be quantified. Literature has shown that there is a distinct relationship between instability number and breakthrough recovery9,10. Table 2 above also includes the instability number associated with each experiment. According to instability theory, while all other parameters remain the same, the lowest injection rate should be the most stable, thus it should have the highest
4

Effect of viscous forces


To investigate the effect of viscous forces two parameters were varied, namely oil viscosity and injection rate. Two oils of different viscosity were used in the experiments. For each oil various injection rates were employed. The properties of the sandpacks are shown in Table 1. These numbers show that the sandpacks have fairly similar characteristics, except for the permeability. In the first two experiments permeability was measured using a pump and pressure transducers. The pressure values are not very accurate for measuring this range of high permeability, so the permeability of the remaining experiments was measured using a constant head pressure. Their values are more accurate, thus it can be assumed that permeability in all experiments was in the range of 2.4 to 2.7 D. The first four experiments used heavy oil 1 (HO1) with a density of 0.98 g/cm3 and a viscosity of 4650 mPa.s. The remaining four experiments used heavy oil 2 (HO2) with a

breakthrough recovery. The results in Table 2 show that most of the experiments follow this trend, indicating that before breakthrough, the effect of instability and viscous fingering dominates the recovery process. However, in the cases of Qinj = 1 mL/hr (0.021 m/d) for both oils, this smallest injection rate did not yield the largest breakthrough recovery as predicted by instability theory. It is not yet known why the recovery of the lowest injection rate was low. One possibility is that water imbibes ahead of the front to cause early breakthrough, and thus lowering the breakthrough recovery. If this is the mechanism responsible for the lower breakthrough recovery, it implies that at ultra-low injection rates, capillary forces once again become important, even for high viscosity oil. Upon careful inspection of the waterflooding response it was seen that even though the breakthrough recovery of the low injection rate is low at the beginning, as more water is injected, the efficiency of the waterflood is significantly better than the other cases, as shown in Figure 3 and Figure 4. These figures show the waterflood response (recovery and water cut) after 5 PVs of brine injected at various rates for HO1. The water-cut profiles show that the water-cut increases drastically as soon as water was injected. They also show that after breakthrough the water cuts are very high, with the higher injection rate having the larger water cut. At first glance, it may appear that the recovery profiles in Figure 3 are very similar at different injection rates, and any differences are merely a reflection of repeatability between experiments. However, Figure 5, which plots the recovery profile for a repeat experiment, shows that this is in fact not the case. The repeatability of these experiments is very high, thus the differences observed for varying injection rates are a reflection of a physical difference in the waterflood responses. The oil recovery behavior of the case where the injection rate is 1 mL/hr is quite interesting. For other cases of higher injection rate, as more water is injected the improvement in recovery plateaus, which implies that the waterflood become less efficient. However, in the case of injecting at 1 mL/hr, oil is continuously produced. Thus proving it is the most efficient, with more oil recovered per pore volume injected, as seen from Table 3. All the numbers in Table 3 are calculated at the later portions of the flood after breakthrough (between 4 and 5 PVs injected). The results in Table 3 shows that the largest injection rate is the least efficient, due to its smallest slope of amount of oil recovered per pore volume injected. Also, these numbers indicate that the experiments performed with HO1 generally were more efficient than for HO2. Again, this is a reflection of HO2s higher viscosity. If viscous force is the dominant recovery mechanism, then as the injection rate increases, the oil production rate should also increase accordingly, as expected from Darcys Law. The fact that this did not happen gives an indication that the recovery mechanism after breakthrough is not solely proportional to viscous forces. This is shown further in Figure 6, which plots the waterflooding recoveries after 5 PVs of brine have been injected against the injection rate. At any particular injection rate, recovery is less for the more viscous oil. For each specific oil, as the injection rate is increased, the recovery after 5 PVs decreased. Since viscous force is proportional to the rate of injection, Figure 6 has shown that as the viscous force increases, total recovery decreases. The results also show that viscous forces are detrimental to recovery not only until breakthrough but even after that when
5

water channels through the porous medium. It was also seen that after breakthrough, recovery of oil is better in the slow injection case. The results from the experiments of different injection rates and oil viscosity values are perhaps not altogether surprising. It is commonly accepted that in situations of poor mobility ratio, severe viscous fingering occurs and this is reflected in lower flood efficiency. These expectations are verified by this data. However, several observations do merit further discussion. One noteworthy observation is the response of the very slow injection rate (1 mL/hr). Contrary to instability theory, this flood showed a lower breakthrough efficiency than the other experiments. At higher injection rates, the growth of viscous fingers leads to instabilities and early water breakthrough. At this lower flow rate, therefore, another competing process is responsible for the flood front being less stable. The second important issue that must be addressed is the mechanism responsible for oil recovery after breakthrough. An interesting analogy for this work is in layered rock systems25, where it has been shown that after water breakthrough any injected water will simply flow through the continuous water channels, and overall recovery is low. In the case of the heavy oil waterfloods, oil is still recovered continuously, despite the high water cut in the produced liquids. Theories of viscous fingering were constructed to explain the recovery efficiency up to breakthrough9-12, therefore although viscous fingers still exist after breakthrough their effect cannot be readily quantified. If oil is still being produced due to viscous flow (i.e. the pressure gradient still existing in the core) then with more pressure (higher injection rate) more oil should be produced. In fact, the opposite happens, which indicates that another mechanism aside from viscous forces is primary responsible for oil production after water breakthrough.

Effect of capillary forces


The effect of capillary forces was first identified when the system of experiment #4 was shut in after 5 PVs of brine had been injected, in order to analyze the results due to its optimistic continuous oil production. After the shut-in period water was injected again at the same rate, immediately the oil fraction in the produced stream increased, as shown in Figure 7. The line in this figure signifies where the shut-in period occurred. The change in slope after this point shows the improvement in recovery. There can only be one possible mechanism that leads to this improved recovery: during the shut-in period, fluid redistribution must have occurred, in which water imbibed into the small pores and displaced oil into the larger pores that had previously been flooded with water. This increase in oil production was seen in all instances when the sandpack was shut-in24. It is therefore established that if imbibition occurred while the sandpack was shut-in, capillary forces are significant even in high viscosity oil systems, in the absence of flow. The question remains is whether imbibition occurs when water is injected continuously. The simple answer is if imbibition occurs during shut-in period, it should occur as well during the injection period, provided that no wettability alteration has occurred. The results at different flow rates were therefore analyzed in order to determine the significance of these forces during water injection. As mentioned before, imbibition rate is a function of time. Thus in order to observe the presence of imbibition, the oil production rate was plotted against time for the experiments performed with HO1 in Figure 8. This figure shows that at

relatively early time (<100,000 s or 28 hrs) the oil production rate decreased accordingly with time. However, as more brine is injected into the porous medium at the slow rate, the reduction in oil production rate became quite small. The gaps in the figure show the shut-in periods with higher oil production after shut-in. The slow reduction in oil production rate at late times indicates the presence of imbibition aiding in the recovery process. An attempt to quantify the effect of viscous and capillary forces is performed, in which the oil production rate was normalized with respect to the lowest injection rate of 1 mL/hr, i.e. the oil production rate obtained at 15 mL/hr was divided by 15. By normalizing all the oil production rates to a single injection rate, the effect of the viscous component is taken out. For the experiments in which HO1 is used, the oil production rate was normalized and plotted against the pore volume of brine injected in Figure 9. This figure shows that when PV injected is less than 2, the normalized oil production rates in all cases are fairly similar, implying the dominant nature of viscous forces at early time. The fact that normalized oil rate is constant in all four cases show that while pressure in the system is high, viscous flow occurs. However, as water continues to be injected the data from the higher injection rates results in lower normalized oil production rate. The same plot was obtained for HO2 (Figure 10), and it shows the same pattern. It is important to clarify at this point that imbibition is still occurring even at higher injection rates. The fact that the normalized rates from the higher injection floods are lower than the actual 1 mL/hr rate simply point to the efficiency of the flood, or the relative influence of the capillary forces. Imbibition is a slow process, especially with heavy oil. At faster injection rates, therefore, there is less time for imbibition to occur per pore volume injected. By reducing the flow rate by a factor of ten, injected fluid now takes ten times as long to traverse the core. Correspondingly, the normalized oil rate is one order of magnitude higher. This is direct evidence that imbibition, which is proportional to time, is responsible for oil production after water breakthrough. In experiment #12, after approximately 7 PVs of brine was injected at 1 mL/hr, the injection rate was increased to 10 mL/hr. After about 41 PVs of brine were injected at 10 mL/hr, the injection rate was then reduced back to 1 mL/hr. Figure 11 shows that increasing the rate of injection improved the oil production rate, which means that viscous forces also cannot be neglected. Even with a reduced effect of capillary forces, increasing the pressure gradient can generate more oil flow. However, for a ten-fold increase in the injection rate, the oil production rate is only doubled. It must also be noted that the water-cut of the products obtained at 10 mL/hr is greater than the case of 1 mL/hr. Thus when evaluating the efficiency of the waterflood, the amount of water produced must be taken into consideration as well. Figure 11 also plots the oil production rates for the fast injection regime, normalized to a low injection rate. This figure further shows that it would be more efficient to inject at a slower rate. These results consistently show that when injecting water at a slow rate, it allows more time for imbibition to contribute to the recovery of oil after breakthrough. The analysis of these experimental findings indicates that at later times (after water breakthrough), oil production is to a large extent due to water imbibition. At slower injection rates, the flood efficiency can be considerably improved, and higher overall recoveries can be obtained.

Prediction of heavy oil waterflooding recovery


It has been shown throughout this study that the waterflooding recovery is a function of various parameters, and that capillary forces can aid in recovery. Thus, it would be beneficial to obtain a general correlation to predict the efficiency of the waterflood. For waterflooding of conventional oil, the performance could be predicted through the capillary number1, which shows that as capillary number increases, recovery increases as well. From Table 2, it was seen that this correlation does not apply to heavy oil due to the severe viscous fingering. Abrams26 tried to take into account the influence of oil viscosity by empirically introducing an extra term in the capillary number: w w cos o

N ca =

0.4

............................................................... (5)

The range of viscosity of oils used in his data set varied from 0.4 to 37 mPa.s. This correlation was applied to the data set used in this study, but the results did not show the trend observed by Abrams. One possible reason for the scatter is that the range of viscosity is much larger than that studied by Abrams, thus a simple correction to capillary number cannot account for severe viscous fingering. As mentioned before, recovery should be related to the instability number. Thus the recoveries of the waterflood are plotted against instability number in Figure 12. Generally, as the instability number increases, recovery decreases. Even though instability theory was only constructed to predict the behavior until breakthrough, this figure indicates that the recovery after breakthrough is somewhat affected by the same factors that control the stability of the flood front. To further explore this relationship results from literature were gathered and plotted along these results on Figure 13. Table 4 summarizes the literature sources and the properties along with the results of the current floods (labeled as A F in Table 4). It must be noted that some of these references did not provide the recovery after 5 PVs of brine injection, thus they were extrapolated to obtain the recovery at 5 PVs. Also, since most of these sources did not report krw at Sor or kro at Swi, these values were approximated, thus the mobility ratio calculated is not accurate. The trend seen in Figure 13 is similar to the one observed with the smaller data set in Figure 12. It was also observed here that the recovery appears to be somewhat constant when the instability number is less than 13.56, which was also found by numerous other researchers9,10,12. However, a significant amount of scattering is present in this relationship. This could be caused by a combination of many factors. As mentioned previously, due to the lack of information some of the data had to be approximated, thus this contributed some errors. Also, it must be kept in mind that these results were gathered from many different researchers, who have used different types of sand, different dimensions and geometry of the sandpacks, as well as different packing techniques. Any of these parameters could have change the value of C* in equation (1). There are other aspects that the instability number did not take into consideration, such as the length or the permeability of the system. The data from Table 4 shows that the range of permeability is from 0.79 D to 95.5 D. Capillary forces are
6

directly related to permeability, thus the data obtained from high vs low permeability systems will have very different responses from water imbibition. This significant variability would definitely contribute to the scattering seen in Figure 13. In order to take these parameters into proper consideration an empirical correlation was derived to predict the recovery of heavy oil waterflood after 5 PVs of brine injected. This expression was designed to be very simple, and to require only information that is commonly obtained during waterflooding. Using data gathered in this work and the compilation of experimental data from the literature, the following equation was obtained: 0.010 k 0.13 0.10 o

heavy oil recovery under waterflooding. Tests were performed for two different heavy oils, at varying water injection rates. The parameters that were investigated were the influence of viscous and capillary forces on oil recovery, and explain the mechanisms for oil recovery after water breakthrough. Viscous forces are observed to be important at early times (under 2 PV of water injected). At very low injection rates (1 mL/hr) the breakthrough recovery actually decreased, which contradicts instability theory. The same results were obtained for both oils and in repeat measurements, thus this is a physical occurrence. Overall, however, oil recovery was observed to decrease with increasing oil viscosity or injection rate. This is an indication of the unstable nature of heavy oil waterfloods. The significance of capillary pressure was investigated, in order to explain the reason for improved oil recovery at low rates. The oil production rates were normalized to the lowest injection rate, which removes the effect of viscous flow. It was then observed that at low injection rates, the normalized oil rate was proportionally higher than the oil flow rates at higher water injection rates. This indicates that capillary forces are significant even during flow of viscous heavy oil. By properly controlling a heavy oil waterflood, the ratio of oil/water flow rates can be improved by as much as one order of magnitude. A secondary goal of this work was to be able to predict the recovery from a heavy oil waterflood. A correlation was observed between recovery and instability number, and a simple empirical model was developed to predict waterflood recovery based on only the injection rate, oil viscosity and sand permeability. This correlation, developed over a wide range of oils and sand permeability, appears to be able to predict the recovery from waterflooding under unstable conditions.

R = 105

.......................................................................... (6)

Equation 6 shows that oil recovery is inversely proportional to absolute permeability (D) and oil viscosity (mPas), and directly proportional to injection velocity (m/s). The effect of permeability and oil viscosity are not unexpected. As permeability decreases (while, of course, still remaining high enough for heavy oil to flow), the effect of imbibition becomes more significant and oil recovery is expected to increase. Likewise, as oil viscosity increases, recovery has been shown to decreases. This is a contribution of both increased instability and also a slower imbibition rate, since water must displace oil in order to imbibe into rock pores. The effect of velocity is somewhat unexpected, given the results of Figure 5, which shows recovery decreasing with increasing injection rate. However, upon closer analysis of the response from viscous vs. capillary forces, the relationship can be explained. It was shown in Figure 9 that in the first two pore volumes injected, oil production was dominated by viscous forces, which are proportional to fluid velocity. Moreover, when fluid was injected at only 1 mL/hr, breakthrough efficiency actually decreased, in contrast to instability theory. It was only after two PV of water had been injected that the recovery became higher than the values at higher injection rates, which had leveled out. Finally, it should be noted that in Figure 10, when the fluid injection rate was increased from 1 10 mL/hr, the oil flow rates did in fact increase accordingly as well. The increase was much smaller than that of the injection rate, however, which was indicative of the low efficiency of injecting at high rates. The fact that the velocity term is in the numerator in Equation 6 is due to the complex nature of the velocity relationship to oil recovery. In terms of velocity, imbibition and viscous flow are actually in competition. The exponent of the velocity term is much smaller than those of permeability and oil viscosity, which reflects the uncertainty in how this parameter affects final recovery. The predicted values are plotted against the actual recoveries in Figure 14. The data used in generating this correlation cover a wide range of injection rates (10-4 10-7 m/s), absolute permeability (0.79 95.5 D) and oil viscosity (12 11500 mPas). This very simple empirical correlation therefore appears to be valid for estimating oil recovery in homogeneous water wet systems of widely varying permeability and oil viscosity.

Acknowledgements
The authors would like to thank Xiao Dong Ji, Jun Gao, John Schnitzler and Jon Bryan of TIPM Laboratory for their contributions. We also wish to thank Nexen Inc. and BP for providing the oil used in these experiments. Financial support of this project was provided by NSERC, COURSE, ISEEE and the Canada Research Chair in Energy and Imaging and its Industrial Affiliates (Shell, Nexen, Devon, PetroCanada, Canadian Natural, ET Energy, Suncor, Schlumberger, Laricina, Paramount).

NOMENCLATURE
C* CT D HO1 HO2 Isr kwor koiw L M Nca PV Qinj R Rbt
7

= = = = = = = = = = = = = = =

Conclusions
A set of ambient temperature laboratory core floods was performed in order to identify the mechanisms responsible for

wettability constant computerized tomography diameter of the core [cm] heavy oil 1 heavy oil 2 Instability number permeability to water at the irreducible oil saturation Sor [D] permeability to oil at the connate water saturation Swi [D] length of sandpack [cm] mobility ratio capillary number pore volume [mL] injection rate [mL/hr] recovery (%OOIP) breakthrough recovery (%OOIP)

t o w 1-D

= = = = = = = =

time [s] contact angle oil viscosity [mPas] viscosity of water [mPas] interfacial tension [mN/m] velocity [m/s] porosity 1-dimensional

17.

18.

REFERENCES
1. 2. 3. GREEN, D.W. and WILHITE, G.P., Enhanced Oil Recovery; SPE Textbook Series Vol. 6, Society of Petroleum Engineers Inc., 1998. MOORE, T.F. and SLOBOD, R.L., The Effect of Viscosity and Capillarity on the Displacement of Oil by Water; Prod. Monthly, 20 30, August 1956. DONG, M. and DULLIEN, F.A.L., Effect of Capillary Forces on Immiscible Displacement in Porous Media; SPE 56676, SPE Annual Technical and Exhibition, Houston, Texas USA, October 3-6, 1999. JENNINGS, H.Y., Waterflood Behaviour of High Viscosity Crudes in Preserved Soft and Unconsolidated Cores; SPE 1202, J. Pet. Tech., 116 120, January 1966. SELBY, R., ALIKHAN, A.A. and FAROUQ ALI, S.M., Potential of non-thermal methods for heavy oil recovery; J. Can. Pet. Tech., Vol. 28, No. 4, July-August 1989. KUMAR, M., HOANG, V. and SATIK, C., High Mobility Ratio Waterflood Performance Prediction: Challenges and New Insights; SPE 97671, SPE International Improved Oil Recovery Conference, Kuala Lumpur, Malaysia, December 5-6 2005. MILLER, K.A., Improving the State of the Art of Western Canadian Heavy Oil Waterflood Technology; J. Can, Pet. Tech., Vol. 45, No. 4, 7 11, 2006. CHATZIS, I., MORROW, N.R. and LIM, H.T., Magnitude and Detailed Structure of Residual Oil Saturation; SPE/DOE 10681, 1982 SPE/DOE 3rd Joint Symposium on Enhanced Oil Recovery, Tulsa, Oklahoma USA, April 4 7 1982. PETERS, E.J. and FLOCK, D.L., The Onset of Instability During Two-Phase Immiscible Displacement in Porous Media; SPE Journal, 249 258, April 1981. BENTSEN, R.G., A New Approach to Instability Theory in Porous Media; SPE Journal, 765 779, October 1985. CHUOKE, R.L., van MEURS, P. and van der POEL, C., Instability of Slow, Immiscible, Viscous Liquid-Liquid Displacements in Permeable Media; Pet. Trans. AIME, 188 194, 216, 1959. SARMA, H.K. and BENTSEN, R.G., An experimental verification of a modified instability theory for immiscible displacements in porous media; J. Can. Pet. Tech., 88 99, July-August 1987. BOBEK, J.E., MATTAX, C.C. AND DENEKAS, M.O., Reservoir Rock Wettability Its Significance and Evaluation; Pet. Trans. AIME, 155 160, 213, 1958. CLARK, K.A., Hot Water Separation of Alberta Bitumenous Sand; Cdn. Inst. Min. & Metall.Trans., Vol. 47, 257 274, 1944. LI, K. and HORNE, R.N., Generalized Scaling Approach for Spontaneous Imbibition: An Analytical Model; SPE Res. Eval. & Eng., 251 258, June 2006. LI, K. and HORNE, R.N., Computation of Capillary Pressure and Global Mobility From Spontaneous Water
8

19. 20.

21.

4.

22.

5. 6.

23. 24.

7. 8.

25.

26.

9. 10. 11.

27.

28.

12.

29.

13. 14. 15. 16.

30.

31.

Imbibition Into Oil-Saturated Rock; SPE Journal, 458 465, December 2005. ZHOU, X., MORROW, N.R. and MA, S., Interrelationship of Wettability, Initial Water Saturation, Aging Time, and Oil Recovery by Spontaneous Imbibition and Waterflooding; SPE Journal, 199 207, June 2000. FISCHER, H. and MORROW, N.R., Modeling the Effect of Viscosity Ratio on Spontaneous Imbibition; SPE 102641, SPE Annual Technical Conference and Exhibition, San Antonio, Texas, September 24 27, 2006. BLAIR, P.M., Calculation of Oil Displacement by Countercurrent Water Imbibition; SPE Journal, 195 202, September 1964. LI, K., CHOW, K. and HORNE, R.N., Effect of Initial Water Saturation on Spontaneous Water Imbibition; SPE 76727, SPE Western Regional/AAPG Pacific Section Joint Meeting, Alaska, May 20 22, 2002. REIS, J.C. and CIL, M., A model for oil expulsion by counter-current water imbibition in rocks: onedimensional geometry; Journal of Petroleum Science and Engineering, 97 107, 10, 1993. NEWCOMBE, J., McGEE, J., RZASA, and M.J., Wettability versus Displacement in Waterflooding in Unconsolidated Sand Columns; Pet. Trans. AIME, 227 232, 204, 1955. MUNGAN, N., Role of Wettability and Interfacial Tension in Water Flooding; SPE Journal, 115 123, June 1964. MAI, A., BRYAN, J., GOODARZI, N. and KANTZAS, A., Insights Into Non-Thermal Recovery of Heavy Oil; WHOC Paper 2006-553, 1st World Heavy Oil Conference, Beijing, CHINA, November 12 15, 2006. HOU, J.R., DONG, M.Z., YANG, J.Z., LIU, Z.C. and YUE, X.A., Effect of Viscosity of Alkaline/Surfactant/Polymer (ASP) Solution on Enhanced Oil Recovery in Heterogeneous Reservoirs; J. Can. Pet. Tech., 27 33, November 2006. ABRAMS, A., The Influence of Fluid Viscosity, Interfacial Tension, and Flow Velocity on Residual Oil Saturation Left by Waterflood; SPE Journal, 437 447, October 1975. SARMA, H.K., MAINI, B.B. and ALLEN, G., Effects of Viscous Instability on Unsteady-State Permeability; CIM/SPE 90-66, International Technical Meeting, Calgary, Alberta, June 10 13, 1990. LIU, Q., DONG, M. and MA, S., Alkaline/Surfactant Flood Potential in Western Canadian Heavy Oil Reservoirs; SPE 99791, SPE/DOE Symposium on Improved Oil Recovery, Tulsa, Oklahoma, April 22 26, 2006. SARMA, H.K., MAINI, B.B., PURVES, R.W. and JHA, K.N., A Laboratory Investigation of the Pseudo Relative Permeability Characteristics of Unstable Immiscible Displacements; J. Can. Pet. Tech., 42 49, January 1994. MAINI, B.B. COSKUNER, G. and JHA, K., A comparison of steady-state and unsteady-state relative permeabilities of viscous oil and water in Ottawa sand; J. Can. Pet. Tech., 72 77, March - April 1990. ZHANG, Y.P., SAYEGH, S. and HUANG, S., Enhanced Heavy Oil Recovery by Immiscible WAG Injection; CIPC 2006-014, Petroleum Societys 7th Canadian International Petroleum Conference, Calgary, June 13 15, 2006.

32. FAROUQ ALI, S.M., FIGUEROA, J.M., AZUAJE, E.A. and FARQUHARSON, R.G., Recovery of Lloydminster and Morichal crudes by caustic, acid and emulsion floods; J. Can. Pet. Tech., 53 59, JanuaryMarch 1979. 33. THOMAS, S., FAROUQ ALI, S.M., SCOULAR, J.R. and VERKOCZY, Chemical Methods for Heavy Oil Recovery; J. Can. Pet. Tech., 56 61, March 2001. 34. SYMONDS, R.W.P., FAROUQ ALI, S.M. and THOMAS, S., A laboratory study of caustic flooding for two Alberta crude oils; J. Can. Pet. Tech., 44 49, January February 1991. 35. MAINI, B.B. and OKAZAWA, T., Effects of temperature on heavy oil-water relative permeability of sand; J. Can. Pet. Tech., 33 41, May June 1987.

Table 4. Summary of results from literature


Source D (cm) L (cm) (fraction) k (D) (mPas) A B C D E F Ref 27 8.89 3.63 3.63 2.6 2.54 2.54 5.64 5.64 5.64 5.64 5.64 5.64 Ref 28 4.25 4.25 114.8 21.1 21.1 53.2 9.3 9.27 46.6 46.6 46.6 46.6 46.6 46.6 14.2 14.2 50.5 50.5 50.5 46.7 46.7 30 30.48 30.48 61 100 60 30 43.7 43.7 43.7 43.7 43.7 0.348 0.450 0.450 0.410 0.460 0.457 0.348 0.348 0.348 0.348 0.348 0.348 0.356 0.358 0.390 0.390 0.390 0.350 0.350 0.391 0.394 0.398 0.374 0.338 0.387 0.236 0.344 0.344 0.344 0.344 0.344 3.86 9.7 9.05 0.79 10 10 3.6 3.6 3.6 3.6 3.6 3.6 6.5 7.1 95.5 95.5 95.5 3.6 3.6 1.3 3.807 1.689 5.11 17.8 9.2 1.039 3 3 3 3 3 11500 11500 11500 11500 11500 11500 10.4 10.6 19 112 110 95 430 1450 717 717 717 450 12 2140 1649 1600 2045 408.3 18 51.2 1190 116 19.4 5.7 2.21 (m/s) 1.23E-07 2.44E-07 4.39E-06 3.04E-06 2.35E-07 4.88E-06 5.92E-05 6.40E-05 1.32E-04 1.32E-04 1.08E-04 2.12E-04 1.11E-06 1.96E-06 2.74E-07 5.47E-07 1.09E-06 1.11E-07 8.89E-07 4.87E-06 7.06E-06 7.06E-06 9.54E-05 5.56E-06 9.26E-05 1.16E-05 1.01E-05 9.88E-06 1.02E-05 1.04E-05 1.08E-05 R (%OOIP) 24 26 20 27 27 21 71.02 69.71 61.13 54.52 54 50.63 51.4 43.1 26.68 21.34 19.21 43.38 55.88 35 35 48 35 53.9 60 50.8 44.79 56.27 54.59 48.99 61.59

Exp # 1 2 3 4 7 8 9 12

Table 1. Properties of sandpacks PV OOIP L (cm) (cm3) f (fraction) k (D) Swi (%) (mL) 16.5 69.99 0.372 1.86 12.25 61.42 17.4 71.71 0.361 1.79 12.77 62.55 16.7 69.40 0.364 2.8 11.65 61.31 16.6 68.00 0.359 2.43 11.62 60.10 17.2 71.75 0.366 2.76 10.33 64.34 16.95 68.70 0.356 2.79 10.61 61.41 17.55 71.01 0.355 2.79 10.34 63.67 16.95 69.16 0.358 2.58 11.17 61.44

Ref 29 *20.3x1 *20.3x1 *20.3x1 Ref 30 Ref 31 Ref 32 Ref 33 Ref 34 5.64 5.64 5 5.08 5.08 5 5.04 5.04 Ref 25 *4.5x4.5 Ref 35 5.64 5.64 5.64 5.64 5.64

Table 2. Recovery and dimensionless numbers Experiment Qinj Rbt Nca Isr (mL/hr) (%OOIP) 1 10 13.82 8.122E-8 905 2 15 11.94 1.218E-7 1234 3 5 15.72 4.061E-8 529 4 1 10.66 8.122E-9 118 7 20 10.57 1.624E-7 5976 8 10 12.03 8.122E-8 2988 9 1 9.97 8.122E-9 299 12 1 and 10 9.42 8.122E-9 299 Table 3. Recovery per pore volume injected Experiment Recovery per PV inj 1 1.61 2 1.27 3 1.31 4 3.71 7 0.88 8 1.52 9 3.21 12 3.31

100

Oil recovery (%OOIP)

Water cut (%)

70 60 50 40 30 20 10 0 0 2 4 PV injected 6 8

95 90 85 80 75 70 0 1 2 3 4 5 PV injected 10 mL/hr 15 mL/hr 5 mL/hr 1 mL/hr

Figure 4. Water-cuts as a function of rate for HO1 Figure 1. Heavy oil waterflooding recovery profile
40
100 90 Water cut (%) 80 70 60 50 40 0 2 PV injected 4 6

Recovery (%OOIP)

30 20 10 0 0 0.5 1 1.5 2 Pore volume injected

Figure 5. Repeatability of heavy oil waterflooding Figure 2. Produced water-cuts in heavy oil waterflood
45 40 35 30 25 20 15 10 5 0 0 5 10 15 20 Qinj (mL/hr) Recovery (%OOIP)

50 Oil recovery (%OOIP) 40 30 10 mL/hr 20 10 0 0 1 2 3 PV injected 15 mL/hr 5 mL/hr 1 mL/hr 4 5

HO1 HO2 25

Figure 6. Recovery as a function of injection rate two oils Figure 3. Waterflooding recoveries at different rates for HO1

10

70 Oil recovery (%OOIP) 60 50 40 30 20 10 0 0 5 PV injected 10 15


Qo normalized (mL/s)

1.E-03

20 mL/hr 10 mL/hr 1 mL/hr 12, 1 mL/hr

1.E-04

1.E-05

1.E-06 0 1 2 3 4 5 6 7 PV injected

Figure 7. Recovery profile before and after fluid redistribution


10 mL/hr

Figure 10. Normalized oil production rates for HO2


20 mL/hr 1.E-03 Qo normalized (mL/s) 1.E-04

1.E-02 1.E-03 Qo (mL/s) 1.E-04 1.E-05 1.E-06 1.E+03 1.E+04 1.E+05 Time (s) 1.E+06

10 mL/hr 1 mL/hr 12, unnormalized 12 normalized

15 mL/hr 5 mL/hr 1 mL/hr

1.E-05 1.E-06 1.E-07

1.E+07

10

20

30 PV injected

40

50

60

Figure 11. Effect of increased viscous forces on oil rates Figure 8. Oil production rates for different water injection rates
10 mL/hr 15 mL/hr 5 mL/hr 1.E-04 1 mL/hr

1.E-03 Qo normalized (mL/s)

1.E-05

45 40 35 30 25 20 15 10 5 0 1 10 100 1000 Instability number

Recovery (% OOIP)

HO1 HO2 10000

1.E-06 0 5 PV injected 10 15

Figure 12. Recovery as a function of instability number Figure 9. Normalized oil production rates for HO1

11

70 Recovery (% OOIP) 60 50 40 30 20 10 0 0.1 10 1000 100000 Instability number

Figure 13. Core floods from different researchers correlation between instability number and recovery

Actual Recovery (%OOIP)

80 60 40 20 0 0 20 40 60 80 Predicted Recovery (%OOIP)

Figure 14. Empirical correlation prediction of oil recovery after 5 PV injected

12

También podría gustarte