Design and Performance of a Novel Autonomous Inflow Control

Nov 30, 2017 - Long horizontal wells have significant advantages in offshore oil and gas development. .... To better-understand the rules of the oil, ...
4 downloads 8 Views 4MB Size
Article pubs.acs.org/EF

Cite This: Energy Fuels XXXX, XXX, XXX−XXX

Design and Performance of a Novel Autonomous Inflow Control Device Lin Zhao,†,‡ Quanshu Zeng,†,‡ and Zhiming Wang*,†,‡ †

State Key Laboratory of Petroleum Resources and Prospecting, College of Petroleum Engineering and ‡MOE Key Laboratory of Petroleum Engineering, College of Petroleum Engineering, China University of Petroleum, Beijing 102249, China ABSTRACT: Due to geological complexity and the heel−toe effect, the production profiles of long horizontal wells are usually imbalanced, and as a result, premature water breakthrough is usually encountered. Once water breakthrough occurs, this phenomenon will reduce oil production. To maximize oil recovery, inflow-control devices (ICDs) are widely used to create a uniform inflow profile. To date, known ICDs cannot meet all the ideal requirements throughout the well’s life. In this study, based on the combination of a successive restriction mechanism and water swelling rubber, a novel autonomous inflow control device is proposed. Then, the rules of oil−water two-phase flow through the novel design are studied by a numerical simulation based on optimized structural parameters, and the fluid property sensitivities are analyzed. Upon integration of the novel design into the test apparatus, flow tests are conducted. The influences of water content, inflow rate, and injection rate on the pressure drop are further analyzed to provide a guide to completion parameter optimization. The results demonstrate that the novel design has a simple structure. Its flow-resistance rating can be easily adjusted. Additionally, the design provides a significant oil and water resistance difference. The pressure-drop ratio of the water relative to oil can be up to 40. The design has a large crosssectional flow area, providing high plugging resistance and high erosion resistance. Moreover, the design is not sensitive to flow rate or oil properties and has a wide application range. Reservoir heterogeneity should be given more attention than the heel−toe effect when optimizing the completion parameters. optimization, and fluid sensitivity analysis; and section 4 describes the implementation and performance testing of the novel design.

1. INTRODUCTION Long horizontal wells have significant advantages in offshore oil and gas development. However, because of the heel−toe effect, the production profiles of long horizontal wells are usually imbalanced, and as a result, water breakthrough is always encountered. This imbalanced phenomenon could be further aggravated if the reservoir is strongly heterogeneous or fractures are developed. By generating an additional pressure loss, inflowcontrol devices (ICDs) facilitate inflow profile uniformity.1−6 Currently, various ICDs have been developed in the industry.7,8 The ICDs use three types of mechanisms to generate an additional pressure drop: the restriction mechanism,9,10 the friction mechanism,11,12 and the incorporation of both mechanisms mentioned above.13−15 Unfortunately, once water coning occurs, the ICD will promote water production, not restrict it. To overcome this problem, a new generation of ICDs is being developed.16 These devices will generate a much-higher flow resistance once water coning occurs and, thus, are referred to as autonomous inflow control devices (AICDs). The movable disk type17−20 uses the balance between the dynamic and static pressures to control the position of a movable disk; however, the disk may be damaged by the pressure difference exerted on it. The passageway type21−25 uses the balance between the inertial and viscous forces to change the flow path; however, the oil viscosity and flow rate range available are quite narrow for each specific design. Shortages in current AICDs limit the application of this technology. This paper describes in detail the design, implementation, and performance of a novel autonomous inflow control device. The remainder of this paper is organized as follows: section 2 provides a novel autonomous inflow control device design; section 3 presents its flow analysis, structural parameter © XXXX American Chemical Society

2. NOVEL DESIGN A novel autonomous inflow control device (Figure 1) is proposed on the basis of the combination of the successive

Figure 1. Structural diagram of the AICD design.

restriction mechanism and water swelling rubber (WSR). As can be observed, the novel design incorporates a series of bulkheads, flow slots, and water swelling rubbers. The bulkheads have two flow slots cut at a 180° angular spacing, and each set Received: September 9, 2017 Revised: November 19, 2017 Published: November 30, 2017 A

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

All of the designs had an annulus inlet, a base pipe inlet, and a base pipe outlet. Both inlets were set as the “velocity inlet,” the outlet was set as the “pressure outlet,” and the rest were defaulted as the “wall.” The standard κ−ε or laminar model was chosen according to the Reynolds number. The mixture or VOF model was selected depending on whether the flow was dispersed or stratified. Because ICDs are usually installed horizontally, gravity was also considered. 3.2. Flow Analysis. The contours of static pressure with varying water contents are shown in Figure 3. As noted, a pressure loss will occur as the flow passes through successive chambers or flow slots. The proportion of loss between both parts changes with flow properties. Because WSR swelling is negligible while the fluid viscosity is high in a low-water-content situation, the local resistance created by the slots is much smaller than the frictional resistance. The WSR will swell as the water content increases, which is revealed by a decrease in the cross-sectional area and a significant increase in the local resistance. Although the viscosity of the mixture will first increase then decrease, the frictional loss due to the viscosity change is assumed to be small relative to the local loss due to the minimum flow area. 3.3. Structural Parameter Optimization. To highlight the excellent performance of the novel design, three other designs with the same false rejection rate (FRR) of 0.8 were compared. These designs include the nozzle-based, helical channel, and tube-type designs. The numbers of FRR represent the equivalent pressure drop magnitude expected in the pressure unit of bars when flowing under the following conditions: a fluid density of 998.2 kg/m3, a fluid viscosity of 1.003 cP, and a flow rate of 30 m3/day. Although not all ICDs have the same geometries, the structural parameters that affect the FRR can be summed up in the minimum flow area (restrictive) and flow path length (friction). The relationships between flow resistance ratings and the cross-sectional flow areas of different ICDs are shown in Figure 4. Equiflow AICD have a fixed cross-sectional flow area. Because the novel design uses a continuous restrictive mechanism to generate an additional pressure drop instead of the monopole restrictive mechanism, its cross-sectional flow area is roughly triple that of the nozzle-based ICD at every FRR. Similarly, the helical channel and tube-type ICDs have a crosssectional flow areas that are double and 1.25 times that of a nozzle-based ICD, respectively. For example, the nozzle-based, tube-type, helical channel, and novel-design cross FRRs = 0.8 at 39.01, 49.25, 81.49, and 119.98 mm2, respectively. The larger the cross-sectional flow area is, the slower the fluid will flow through that area. Of all of the ICDs studied, the novel design is most resistant to plugging during mud flow-back operations and fluid-borne particle erosion during production. The relationships between the FRR and flow path length in different ICD designs are shown in Figure 5. The results demonstrate that the flow resistance ratings of all these designs increase linearly with the flow path length. The intercepts at the y-axis mean that no frictional flow resistance exists. The nozzlebased ICD, the tube-type ICD, the helical-channel ICD, the Equiflow AICD, and the novel design will cross the y-axis at the 0.760, 0.527, 0.035, 0.136, and 0.043 (0.171) FRR marks, respectively, for Sw = 0% (Sw = 50%). These results indicate that the nozzle-based ICD mainly depends on a restrictive mechanism, whereas the helical channel ICD and the Equiflow AICD mainly depend on a frictional mechanism. Furthermore, the tube-type ICD depends on both restrictive and frictional

of slots is staggered at a 90° phase with the next set; thus, the flow direction must turn after passing through each set of slots, where the jetting effect is prevented. As previously described, the primary pressure drop mechanism of the novel design is restrictive in a distributive variable configuration. The design is insensitive to oil viscosity. Each set of slots has a comparable contribution to the pressure loss. The flow resistance rating can be easily adjusted by adding or reducing the slot sets. Additionally, the WSR installed on the slot will swell once water breakthrough occurs, and the swell increment will be adjusted automatically according to water content (Figure 2). The water swelling rubber is a dynamic

Figure 2. Diagram of the minimum-flow area for different water contents.

balance, swelling to change the flow area and flow resistance rating. Specifically, for a low water content, the swell increment of the WSR is limited; the corresponding flow area is maximized while the flow resistance rating is at a minimum. As the water content increases, the WSR begins to swell; the flow area decreases, whereas the flow resistance rating increases. For a high water content, once the WSR is fully expanded, its flow area is minimized, while its flow resistance rating is at a maximum. In general, the novel design is advanced as follows. The design is under a simple structure; thus, its flow resistance rating can be easily adjusted. Additionally, the design provides a large crosssectional flow area and generates a significant low fluid velocity, resulting in a high plugging resistance and a high erosion resistance. The design also provides a wide application range because it is not sensitive to flow rate or oil properties. Finally, the design provides a significant oil and water resistance difference because it is especially sensitive to water content. For a typical design proposed in this paper, the pressure-drop ratio of water relative to oil can be up to 40.

3. NUMERICAL SIMULATION AND STRUCTURAL PARAMETER OPTIMIZATION In this section, the rules of oil and water two-phase flow through the novel design are studied by numerical simulation. This guides the following structural parameter optimization and fluid property sensitivity analysis. 3.1. Modeling. Mechanical models with different structural parameters were set up using professional modeling software. Next, hydraulic models were obtained by a Boolean operation. To describe the flow around the slots and ports more accurately, the mesh was divided into several regions and encrypted in these locations. B

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

Figure 3. Contours of the static pressure with varying water contents.

Figure 5. FRR vs flow length. Figure 4. FRR vs minimum-flow area.

oil viscosity sensitivities (Table 1) were investigated. In particular, water property sensitivity is not considered due to its constant property. The relationships between the water content and pressure drop are presented in Figure 6. The water contents (%) are as follows: 0, 10, 20, 30, 40, 50, 60, 70, 80, 90, and 100. The pressure drop values of the three passive designs will first increase and then decrease with an increase in water content, and the boundary point lies in the same location as the phase-inversion point. Unfortunately, the pressure drop generated by water is lower than that of oil or the oil−water mixture in most cases. That is, once water breaks through, the current passive ICDs will be ineffective. For the Equiflow AICD, the pressure drop

mechanisms, and the novel design depends on a continuous restrictive mechanism. The FRR increase in the nozzle-based ICD is the highest because it has the highest maximum flow velocity. This finding further proves that the novel design restricts the oil by frictional resistance while restricting the water by a restrictive mechanism. In addition, the novel design adjusts to different FRRs simply by adding or reducing the number of bulkheads. 3.4. Fluid Property Sensitivity Analysis. To betterunderstand the rules of the oil, water, and oil−water mixture flow through the novel design, the water content, oil density, and C

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels Table 1. Fluid Property Sensitivity Research Projects water content (%)

oil density (kg/m3)

oil viscosity (cP)

10 0, 10, 20, 30, 40, 50, 60, 70, 850 80, 90, 100 0, 50 800, 850, 900, 950, 1 1000 0, 50 850 1, 2, 4, 10, 20, 30, 50, 100, 150, 200

Figure 8. Sensitivity analysis of the oil viscosity.

The pressure drops in all these designs increase linearly with oil viscosity. The fluid viscosity sensitivities of these designs vary greatly, The nozzle-based ICD exhibit an increase of 397.9 Pa/cP in the amplitude; the tube-type ICD shows an increase of 3001.7 Pa/cP; the helical channel ICD exhibit an increase of 6401.2 Pa/cP; and the novel design increases 361.3 (45.71) Pa/cP for Sw = 0% (Sw = 50%). These findings indicate that for the designs tested, the novel design will thus provide the best results because it has the lowest sensitivity to viscosity variations.

Figure 6. Sensitivity analysis of the water content.

generated by water is nearly 8 times that generated by oil. With the help of the novel design, the pressure drop generated by water is nearly 40 times that generated by oil. Although the flow pattern or flow regime may change, the pressure drop in the novel design increases monotonously with water content. The relationships between oil density and pressure drop are presented in Figure 7. Because a typical oil density ranges from

4. PERFORMANCE ANALYSIS 4.1. Implement. The main body of the novel design was processed by stainless steel, and PZ-150 hydrophilic expansion rubber was installed in the slots, as shown in Figure 9. To better-understand the impacts of the fluid properties on the swelling property of PZ-150, swelling tests with varying water contents were conducted, as shown in Figure 10. As can be observed, the expansion ratio become steady if the material has been soaked in the oil−water mixture for 1 h. The material response is quick. The swelling characteristics with varying fluid properties differ greatly. In fact, a phase inversion will occur as the water content increases. The inversion point is 13% water content for the 132 cP oil studied. In the water-in-oil case, the water molecule surrounded by oil molecules can hardly enter the material; thus, the WSR swelling is limited. The oil-in-water situation is just the opposite, and the free water can be easily attached to the WSR, resulting in WSR swelling. The higher the water cut is, the faster the expansion rate will be. 4.2. Test Procedure. The overall performance of the novel design was tested upon an apparatus capable of simulating gas/oil/water multiphase variable mass flow in the production section,26,27 as shown in Figure 11. The wellbore inclination can be adjusted from 0° (vertical) to 90° (horizontal). The horizontal direction was selected in this study. The novel design was integrated into the test section, which is transparent with a diameter of 5.5 in and a length of 2.4 m. Refined mineral oil and urban tap water were used as the working fluids. All of the transducers were connected to the data acquisition and control system to allow the continuous monitoring and collection of pressures and flow rates. A system feedback structure controlled the boundary conditions if any adjustment was required. The environmental pressure and temperature were 1 atm and 25 °C, respectively. Oil was first circulated in the apparatus to ensure its sealing performance. Next, water was added into the tank to obtain the desired water content, and another circulation hour was required to ensure that the WSR was fully expanded.

Figure 7. Sensitivity analysis of the oil density.

800 to 1000 kg/m3, the oil densities (kg/m3) are as follows: 800, 850, 900, 950, and 1000. The results indicate that the pressure drop increases linearly with oil density. The oil density sensitivities of these designs are not the same, i.e., the nozzle-based, tubetype, and helical-channel ICDs exhibit increases in amplitude of 78.0, 47.6, and 18.6 Pa/(kg/m3), respectively, whereas the novel design shows an increase of 29.8 (44.0) Pa/(kg/m3) for Sw = 0% (Sw = 50%). The results indicate that for these designs tested, the nozzle-based ICD is the most sensitive to density variations. However, a greater increase in amplitude generates a lower pressure drop at a lower density, and the nozzle-based ICD will thus will provide the best results in this regard. The novel design also shows good performance in this respect. The relationships between the oil viscosity and pressure drop of different designs are illustrated in Figure 8. Because typical oil viscosities ranges from 1 to 200 cP, the oil viscosities (cP) are as follows: 1, 2, 4, 10, 20, 30, 50, 100, 150, and 200. D

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

Figure 9. Novel autonomous inflow device.

4.3. Projects and Results. To provide guidance regarding the completion parameter optimization, this paper further studies the influence of water content, inflow rate, and injection ratio on the pressure drop (Table 2). The Halliburton EquiFlow AICD was used as a reference. The excellent performance of the AICD depends on its water resistance relative to the oil resistance. The pressure drops with varying water contents are shown in Figure 12. As can be noted, the pressure drops of both designs increase with water content. The pressure-drop difference of the Halliburton EquiFlow AICD can be up to a factor of six. Because the water contents of different sections vary between 10% and 30%, the resulting pressure drop difference is not that significant. Thus, the application of this technology is significantly limited. For the novel design, the water resistance relative to oil resistance can be up to 40. The resistance difference can be quite significant, even the water content difference is small.

Figure 10. Expansion ratio variation at different water contents.

Then, both the main flow rate and inflow rate were adjusted to the target value. Finally, the pressure loss can be estimated by averaging multiple measurements after the system was stable.

Figure 11. Schematic of the multi-phase flow test apparatus. E

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

related to the flow pattern of the mixture. In this case, water cannot be effectively controlled at low-permeability sections. For the novel design, the water resistance relative to the oil resistance does not depend on the flow rate. The ratio of water pressure drops to oil pressure drops ranges from 33 to 37. Water can be effectively controlled in the inflow rate range studied. Because horizontal wells are usually segmented in several sections, the main flow rate at each section of the wellbore may be quite different. To study the influence of the heel−toe effect on completion parameter optimization, the pressure drop with varying main flow rates were investigated, as shown in Figure 15.

Table 2. Research Projects water content (%)

inflow (m3/h)

Q/Qt

0, 5, 10, 12, 15, 18, 20, 30, 40, 50, 60, 70, 0.8 1:1 80, 90, 100 0, 5, 10, 12, 15, 18, 20, 30, 40, 50, 60, 70, 0.2, 0.4, 0.6, 0.8 1:1 80, 90, 100 0, 5, 10, 12, 15, 18, 20, 30, 40, 50, 60, 70, 0.2 1:1, 1:10; 1:20; 80, 90, 100 1:40

Figure 12. AICD pressure drops with varying water contents.

Because a horizontal well completed with ICD is relatively long, the well may cross several formations. The reservoir properties at different formations may be quite different, resulting in inflow differences. To study the influence of reservoir heterogeneity on completion parameter optimization, the pressure drops with varying inflow rates were researched, as shown in Figures 13 and 14. For the Halliburton EquiFlow AICD, the

Figure 15. AICD pressure drops with varying injection ratios.

The results show that the AICD pressure-drop change with the main flow rate is assumed to be small. That is, the effect of the AICD location along the wellbore on completion optimization is not necessary.

5. CONCLUSIONS In this paper, a novel autonomous inflow control device design is proposed. The following conclusions and recommendations can be obtained on the basis of the study. (1) The novel AICD has a simple structure. Its flow resistance rating can be easily adjusted by adding or reducing the slot sets. (2) Compared with the traditional AICDs, the novel design provides a significant oil and water resistance difference. The pressure drop ratio of water relative to oil can be up to 40. (3) The design has a large cross-sectional flow area, providing a high plugging resistance and high erosion resistance. (4) The design is not sensitive to flow rate or oil properties, which enables a wide application range. (5) Reservoir heterogeneity should be given more attention than the heel-toe effect when optimizing the completion parameters.

Figure 13. Pressure drops of the EquiFlow AICD with varying inflow rates.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Phone: +86-1089734958. ORCID

Zhiming Wang: 0000-0002-9301-1942 Notes

The authors declare no competing financial interest.



Figure 14. Pressure drops in the novel design with varying inflow rates.

ACKNOWLEDGMENTS Financial support from the Science Foundation of China University of Petroleum, Beijing (no. 2462017YJRC058), National

greater the inflow rate is, the greater the water pressure drop is relative to the oil pressure drop. However, this ratio is heavily F

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX

Article

Energy & Fuels

of the SPE Middle East Oil and Gas Show and Conference; SPE: Manama, Bahrain, 2013. (18) Mathiesen, V.; Werswick, B.; Aakre, H.; Elseth, G. Autonomous valve, a game changer of inflow control in horizontal wells. In Proceedings of the SPE Offshore Europe Oil & Gas Conference & Exhibition; SPE: Aberdeen, UK, 2011. (19) Mathiesen, V.; Werswick, B.; Aakre, H. The next generation inflow control, the next step to increase oil recovery on the Norwegian Continental Shelf. In Proceedings of the SPE Bergen One Day Seminar; SPE: Bergen, Norway, 2014. (20) Carpenter, C. Optimized design of autonomous inflow-control devices for gas and water coning. JPT, J. Pet. Technol. 2015, 67 (12), 70−71. (21) Least, B.; Greci, S.; Burkey, R. C.; Ufford, A.; Wileman, A. Autonomous ICD single phase testing. In Proceedings of the SPE Annual Technical Conference and Exhibition; SPE: San Antonio, TX, 2012. (22) Least, B.; Greci, S.; Wileman, A.; Ufford, A. Autonomous ICD range 3B single phase testing. In Proceedings of the SPE Annual Technical Conference and Exhibition; SPE: New Orleans, LA, 2013. (23) Zeng, Q. S.; Wang, Z. M.; Wang, X. Q.; Wei, J. G.; Zhang, Q.; Yang, G. A novel autonomous inflow control device design and its performance prediction. J. Pet. Sci. Eng. 2015, 126, 35−47. (24) Zeng, Q. S.; Wang, Z. M.; Wang, X. Q.; Li, Y. W.; Zou, W. L. A novel autonomous inflow control device design: Improvements to hybrid ICD. In Proceedings of the International Petroleum Technology Conference; IPTC: Kuala Lumpur, Malaysia, 2014. (25) Zhao, L.; Least, B.; Greci, S.; Wileman, A. Fluidic diode autonomous ICD range 2A single phase testing. In Proceedings of the SPE Oilfield Water Management Conference and Exhibition; SPE: Kuwait City, Kuwait, 2014. (26) Wang, Z. M.; Zhang, Q.; Zeng, Q. S.; Wei, J. G. A unified model of oil/water two-phase flow in the horizontal wellbore. SPE J. 2017, 22 (01), 353−364. (27) Zeng, Q. S.; Wang, Z. M.; Wang, X. Q.; Zhao, Y. L.; Guo, X. A novel oil-water separator design and its performance prediction. J. Pet. Sci. Eng. 2016, 145, 83−94.

Natural Science Foundation of China (no. 51474225), and National Natural Science Foundation of China (no. 51521063) are gratefully acknowledged.



REFERENCES

(1) Rajesh, K.; Vivek, S.; Surseh, K.; Atul, K. Effect of reservoir heterogeneities and model uncertainties on prediction versus actual field behavior- a case study. In Proceedings of the SPE Oil and Gas India Conference and Exhibition; SPE: Mumbai, India, 2017. (2) Hill, A. D.; Zhu, D. The relative importance of wellbore pressure drop and formation damage in horizontal wells. SPE Prod. & Oper. 2008, 23 (2), 232−240. (3) Tommy, T.; Cindy, J. S.; Georg, M. D.; Sarawak, S. B. Reservoir heterogeneities impact gas water contact movement of a mature carbonate field in central luconia. In Proceedings of the Offshore Technology Conference Asia; OTC: Kuala Lumpur, Malaysia, 2016. (4) Grubert, M. A.; Wan, J.; Ghai, S. S.; Livescu, S.; Brown, W. P.; Long, T. A. Coupled completion and reservoir simulation technology for well performance optimization. In Proceedings of the SPE Annual Technical Conference and Exhibition; SPE: New Orleans, LA, 2009. (5) Livescu, S.; Brown, W. P.; Jain, R.; Grubert, M. A.; Ghai, S. S.; Lee, L. W.; Long, T. A. Application of a coupled wellbore/reservoir simulator to well-performance optimization. In Proceedings of the SPE Annual Technical Conference and Exhibition; SPE: Florence, Italy, 2010. (6) Crow, S. L.; Coronado, M. P.; Mody, R. K. Means for passive inflow control upon gas breakthrough. In Proceedings of the SPE Annual Technical Conference and Exhibition; SPE: San Antonio, TX, 2006. (7) Li, Z.; Fernandes, P. X.; Zhu, D. Understanding the roles of inflow-control devices in optimizing horizontal-well performance. SPE Drill. Completion 2011, 26 (3), 376−385. (8) Carpenter, C. ICD well history and future use in a giant oil field offshore Abu Dhabi. JPT, J. Pet. Technol. 2015, 67 (11), 105−106. (9) Aadnoy, B. S.; Hareland, G. Analysis of inflow control devices. In Proceedings of the SPE Offshore Europe Oil & Gas Conference & Exhibition; SPE: Aberdeen, 2009. (10) Vela, I.; Viloria-gomez, L. A.; Caicedo, R.; Porturas, F. Well production enhancement results with inflow control devices (ICD) completions in horizontal well in Ecuador. In Proceedings of the SPE EUROPEC/EAGE Annual Conference and Exhibition; SPE: Vienna, Austria, 2011. (11) Brekke, K.; Lien, S. C. New and simple completion methods for horizontal wells improve production performance in high-permeability thin oil zones. SPE Drill. Completion 1994, 9 (3), 205−209. (12) Visosky, J. M.; Clem, N. J.; Coronado, M. P.; Peterson, E. R. Examining erosion potential of various inflow control devices to determine duration of performance. In Proceedings of the SPE Annual Technical Conference and Exhibition; SPE: Anaheim, California, 2007. (13) Coronado, M. P.; Garcia, L.; Russell, R. D.; Garcia, G. A.; Peterson, E. R. New inflow control device reduces fluid viscosity sensitivity and maintains erosion resistance. In Proceedings of the Offshore Technology Conference; OTC: Houston, TX, 2009. (14) Garcia, L.; Coronado, M. P.; Russell, R. D.; Garcia, G. A.; Peterson, E. R. The first passive inflow control device that maximizes productivity during every phase of a well’s life. In Proceedings of the International Petroleum Technology Conference; UPTC: Doha, Qatar, 2009. (15) Youl, K. S.; Harkomoyo, H.; Suhana, W.; Regulacion, R. E.; Jorgensen, T. Passive inflow control devices and swellable packers control water production in fractured carbonate reservoir: A comparison with slotted liner completions. In Proceedings of the SPE/IADC Drilling Conference and Exhibition; SPE: Amsterdam, The Netherlands, 2011. (16) Faisal, T. A.; Vasily, M. B.; Michael, R. K.; David, D. Advanced wells: a comprehensive approach to the selection between passive and active inflow-control completions. SPE Pro. & Oper. 2010, 25 (3), 305−325. (17) Aakre, H.; Halvorsen, B.; Werswick, B.; Mathiesen, V. Smart well with autonomous inflow control valve technology. In Proceedings G

DOI: 10.1021/acs.energyfuels.7b02673 Energy Fuels XXXX, XXX, XXX−XXX