Skip to main content
Log in

Impact of Two-Phase Flow Pattern on Solvent Vapour Extraction

  • Published:
Transport in Porous Media Aims and scope Submit manuscript

A Correction to this article was published on 27 October 2023

This article has been updated

Abstract

Solvent vapour extraction (Vapex) is a promising technology for in-situ heavy oil recovery from oil sands deposits. The prediction of oil recovery rates requires fundamental understanding of the pore-scale mechanisms and their impact on mass transfer and oil production. To bridge the gap between pore-scale mechanisms and macro-scale recovery, a dynamic pore-network model for two-phase flow with mass transfer is developed. The impact of pressure gradient on two-phase flow pattern, mass transfer and oil production are investigated. It is found that at high capillary numbers, in viscous dominated flow, dissolved oil is moved in intermittent liquid clusters to the outlet of the network. This mechanism of interface renewal maintains a steep solvent mole fraction gradient at the interface and enhances mass transfer, resulting in high oil production. In capillary-dominated flow, capillary fingering with low mass transfer and oil production are observed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Data Availability

Videos of the variation of the distribution of liquid saturation and solvent mole fraction in the liquid phase can be found in the supplementary material.

Change history

References

  • Aghaei, A., Piri, M.: Direct pore-to-core up-scaling of displacement processes: dynamic pore network modeling and experimentation. J. Hydrol. 522, 488–509 (2015)

    Article  Google Scholar 

  • An, S., Erfani, H., Godinez-Brizuela, O. E., Niasar, V.: Transition from viscous fingering to capillary fingering: Application of GPU-based fully implicit dynamic pore network modeling. Water Resour. Res. 56, e2020WR028149 (2020). https://doi.org/10.1029/2020WR028149.

  • Asok Kumar, T.: Measurements of molecular diffusion coefficients of carbon dioxide, methane and propane in heavy oil under reservoir conditions. Master of Applied Science, University of Regina (2004)

  • Bayestehparvin, B., Farouq Ali, S.M., Kariznovi, M., Abedi, J.: Pore-scale modelling of solvent-based recovery processes. Energies 14(5), 1405 (2021)

    Article  Google Scholar 

  • Mokrys, I.J., Butler, R.M.: The rise of interfering solvent chambers: solvent analog model of steam assisted gravity drainage. J. Can. Pet. Technol. 32, 26–36 (1993)

    Article  Google Scholar 

  • Butler, R.M., Mokrys, I.J.: Solvent analog model of steam-assisted gravity drainage. AOSTRA J. Res. 5, 17–32 (1989)

    Google Scholar 

  • Chatzis, I.: Pore scale phenomena of heavy oil recovery using vapor extraction. SCA2002-26 (2002)

  • Chen, S., Qin, C., & Guo, B.: Fully implicit dynamic pore-network modeling of two-phase flow and phase change in porous media. Water Resour. Res. 56, e2020WR028510 (2020). https://doi.org/10.1029/2020WR028510

  • Cuthiell, D., Edmunds, N.: Thoughts on simulating the Vapex process. SPE-158499-MS (2012)

  • Das, S.K., Butler, R.M.: Extraction of heavy oil and bitumen using vaporized hydrocarbon solvents. Pet. Sci. Technol. 15(1–2), 51–75 (1997)

    Article  Google Scholar 

  • Das, S.K., Butler, R.M.: Mechanism of the vapor extraction process for heavy oil and bitumen. J. Pet. Sci. Eng. 21, 43–59 (1998)

    Article  Google Scholar 

  • Haynes, H.W.: A note on diffusive mass transport. Chem. Eng. Educ. 20, 22–27 (1986)

    Google Scholar 

  • James, L., Chatzis, I.: Details of gravity drainage of heavy oil during vapor extraction. In: Proceedings of the International Symposium of the Society of Core Analysts, Abu Dhabi, United Arab Emirates, Society of Core Analysts (2004)

  • Joekar-Niasar, V.: Pore-scale modelling techniques: balancing efficiency, performance, and robustness. Comput. Geosci. 20(4), 773–775 (2016)

    Article  Google Scholar 

  • Joekar-Niasar, V., Hassanizadeh, S.M., Dahle, H.K.: Non-equilibrium effects in capillarity and interfacial area in two-phase flow: dynamic pore-network modelling. J. Fluid Mech. 655, 38–71 (2010)

    Article  Google Scholar 

  • Knorr, K.D., Imran, M.: Extension of Das and Butler semianalytical flow model. J. Can. Pet. Technol. 50(06), 53–60 (2011)

    Article  Google Scholar 

  • Lenormand, R., Touboul, E., Zarcone, C.: Numerical models and experiments on immiscible displacements in porous media. J. Fluid Mech. 189, 165–187 (1988)

    Article  Google Scholar 

  • Li, X., Berg, S., Castellanos-Diaz, O., Wiegmann, A., Verlaan, M.: Solvent-dependent recovery characteristic and asphaltene deposition during solvent extraction of heavy oil. Fuel 263, 116716 (2020)

    Article  Google Scholar 

  • Luo P., Kladkaew, N., Gu, Y.: Detection of the onset of asphaltene precipitation in a heavy oil-solvent system. Paper 2007-096 (2007)

  • Martinez, J.F., Schoeggl, F.F., Maini, B.B., Yarranton, H.W.: Investigation of mechanisms for gravity drainage of heavy oil and solvent mixtures in a Hele-Shaw cell. Energy Fuels 34(5), 5823–5837 (2020)

    Article  Google Scholar 

  • Moghadam, S., Nobakht, M., Gu, Y.: Theoretical and physical modeling of a solvent vapour extraction (VAPEX) process for heavy oil recovery. J. Pet. Sci. Eng. 65(1–2), 93–104 (2009)

    Article  Google Scholar 

  • Nenniger, J.E., Dunn, S.G.: How fast is solvent based gravity drainage? SPE-2008-139 (2008)

  • NIST. https://webbook.nist.gov/chemistry/fluid/

  • Patzek, T.W.: Verification of a complete pore network simulator of drainage and imbibition. SPE 71310, 144–156 (2000)

    Google Scholar 

  • Qi, Z., Abedini, A., Lele, P., Mosavat, N., Guerrero, A., Sinton, D.: Pore-scale analysis of condensing solvent bitumen extraction. Fuel 193, 284–293 (2017)

    Article  Google Scholar 

  • Rezaei, N., Mohammadzadeh, O., Parsaei, R., Chatzis, I.: The effect of reservoir wettability on the production characteristics of the VAPEX process: an experimental study. Transp. Porous Media 90(3), 847–867 (2011)

    Article  Google Scholar 

  • Shi, Y., Yang, D.: Experimental and theoretical quantification of nonequilibrium phase behavior and physical properties of foamy oil under reservoir conditions. J. Energy Resour. Technol. 139(6), 062902 (2017)

    Article  Google Scholar 

  • Shu, W.R.: A viscosity correlation for mixtures of heavy oil bitumen and petroleum fractions. Soc. Pet. Eng. J. 24, 277–282 (1984)

    Article  Google Scholar 

  • Soiket, M.I.H., Oni, A.O., Kumar, A.: The development of a process simulation model for energy consumption and greenhouse gas emissions of a vapor solvent-based oil sands extraction and recovery process. Energy 173, 799–808 (2019)

    Article  Google Scholar 

  • Stewart, R.A., Shaw, J.M.: Interface renewal and concentration shock through sloughing: accounting for the dissonance between production models and measured outcomes for solvent-assisted bitumen-production processes. SPE-186108-PA (2018)

  • Talebi, S., Abedini, A., Lele, P., Guerrero, A., Sinton, D.: Microfluidics-based measurement of solubility and diffusion coefficient of propane in bitumen. Fuel 210, 23–31 (2017)

    Article  Google Scholar 

  • Toro Monsalve, I.C., Oni, A.O., Kumar, A.: Development of techno-economic models for assessment of solvent-based bitumen extraction technology. Energy Convers. Manag. 253, 115142 (2022)

    Article  Google Scholar 

  • Upreti, S.R., Lohi, A., Kapadia, R.A., El-Haj, R.: Vapor extraction of heavy oil and bitumen: a review. Energy Fuels 21, 1562–1574 (2007)

    Article  Google Scholar 

  • Vidales, A.M., Riccardo, J.L., Zgrablich, G.: Pore-level modelling of wetting on correlated porous media. J. Phys. D Appl. Phys. 31, 2861–2868 (1998)

    Article  Google Scholar 

  • Wang, Q., Chen, Z.: Transient mass transfer ahead of a hot solvent chamber in a heavy oil gravity drainage process. Fuel 232, 165–177 (2018)

    Article  Google Scholar 

  • Wang, Q., Jia, X., Chen, Z.: Mathematical modeling of the solvent chamber evolution in a vapor extraction heavy oil recovery process. Fuel 186, 339–349 (2016)

    Article  Google Scholar 

  • Wang, Q., Jia, X., Chen, Z.: Modelling of dynamic mass transfer in a vapour extraction heavy oil recovery process. Can. J. Chem. Eng. 95(6), 1171–1180 (2017)

    Article  Google Scholar 

  • Wang, H., Shaw, J.M., Jin, Z.: Discontinuous displacement at solvent-immobile hydrocarbon interfaces. Energy Fuels 34(8), 9392–9400 (2020)

    Article  Google Scholar 

  • Yang C., Gu, Y.: Effects of heavy-oil/solvent interfacial tension on gravity drainage in the vapor extraction (Vapex) process. SPE-97906-MS (2005)

  • Yang, C., Gu, Y.: Diffusion coefficients and oil swelling factors of carbon dioxide, methane, ethane, propane, and their mixtures in heavy oil. Fluid Phase Equilib. 243(1–2), 64–73 (2006)

    Article  Google Scholar 

Download references

Funding

The authors acknowledge support from the Government of Canada’s Interdepartmental Program of Energy Research and Development (PERD).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Merouane Khammar.

Ethics declarations

Competing interests

The authors declare that they have no financial interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The original online version of this article was revised: Copyright line updated.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary file1 (MP4 11351 kb)

Supplementary file2 (MP4 14735 kb)

Supplementary file3 (MP4 7900 kb)

Supplementary file4 (MP4 9233 kb)

Supplementary file5 (MP4 5282 kb)

Supplementary file6 (MP4 6303 kb)

Supplementary file7 (MP4 6602 kb)

Supplementary file8 (MP4 10165 kb)

Supplementary file9 (MP4 6525 kb)

Supplementary file10 (MP4 7783 kb)

Supplementary file11 (MP4 5345 kb)

Supplementary file12 (MP4 10748 kb)

Supplementary file13 (DOCX 10759 kb)

Appendices

Appendix A: Local Rules

The analytical expressions used for the calculation of capillary pressures in pores, hydraulic conductivities for liquid and gas phases and conditions for throat invasion and snap-off are presented in this appendix.

1.1 A.1 Pore Capillary Pressure

Pore bodies are modelled as cubes with the wetting phase residing in the corners and along the edges. The Capillary pressure in a pore body, \({p}_{i}^{c}\), depends on the inscribed radius, Ri, and liquid saturation, \({S}_{i}^{l}\), following the equation (Joekar-Niasar et al. 2010):

$$p_{i}^{c} \left( {s_{i}^{l} } \right) = \frac{2 \sigma }{{R_{i} \left( {1 - \exp \left( { - 6.83 s_{i}^{l} } \right)} \right)}},$$
(A.1)

where \(\sigma\) is the interfacial tension between wetting and non-wetting fluids.

A minimum liquid-phase saturation is associated with each pore body, \(s_{{i, {\text{min}}}}^{l}\). It depends on the global pressure difference imposed between the inlet and outlet of the model, \(p_{{{\text{global}}}}^{c}\). It is given by the following equation:

$$s_{{i,{\text{min}}}}^{l} = - \frac{1}{6.83}\;\ln \left( {1 - \frac{1}{{R_{i} }}\frac{2 \sigma }{{p_{{{\text{global}}}}^{c} }}} \right).$$
(A.2)

1.2 A.2 Throat Invasion

The vapour phase invades a pore throat when the capillary pressure in the neighbouring pore body exceeds the entry capillary pressure of the pore throat. It is given by the following equation (Joekar-Niasar et al. 2010):

$$p_{e, ij}^{c} = \frac{\sigma }{{r_{ij} }} \left( {\cos \theta + \sqrt {\frac{\pi }{4} - \theta + \sin \theta \cos \theta } } \right),$$
(A.3)

where \(r_{ij}\) is the radius of the inscribed circle of the pore throat cross section, and \(\theta\) is the contact angle.

1.3 A.3 Snap-Off

When the capillary pressure in a pore throat drops below a critical value, liquid in the corners snaps-off and fills the throat. For a throat with square cross section, snap-off occurs when (Vidales et al. 1998):

$$p_{ij}^{c} \le \frac{\sigma }{{r_{ij} }} \left( {\cos \theta - \sin \theta } \right).$$
(A.4)

1.4 A.4 Throat Conductance

When the throat is occupied by a single phase \(\alpha\), phase conductance is given by the following equation (Patzek 2000):

$$K_{ij}^{\alpha } = \frac{{0.5623 G_{ij} \left( {A_{ij} } \right)^{2} }}{{\mu^{\alpha } L_{ij} }},$$
(A.5)

\(G_{ij} = \frac{{A_{ij} }}{{\left( {C_{ij} } \right)^{2} }}\), is the shape factor of the pore throat cross section. \(A_{ij}\), \(C_{ij}\) and \(L_{ij}\) are the cross-sectional area, the perimeter and the length of the throat connecting pores i and j. The length of the throat connecting pores i and j is assumed to be equal the distance between the centre of the pores.

When a pore throat is invaded and two-phases coexist along the throat as shown in the figure above, the conductivity of the phase is given by the following equation:

$$K_{ij}^{g} = \frac{{0.5623 G_{ij}^{g} \left( {A_{ij}^{g} } \right)^{2} }}{{\mu^{g} L_{ij} }},$$
(A.6)

\(G_{ij}^{g} = \frac{{A_{ij}^{g} }}{{\left( {C_{ij}^{g} } \right)^{2} }}\), is the shape factor of the non-wetting phase region in the pore throat cross section; \(A_{ij}^{g}\), is the cross-sectional area of the gas phase region, given by:

$$A_{ij}^{g} = 4 R_{ij}^{2} - 4 r_{ij}^{2} \left[ {\frac{{\sin \left( {\frac{\pi }{4} - \theta } \right)}}{{\sin \left( {\frac{\pi }{4}} \right)}}\cos \left( \theta \right) - \left( {\frac{\pi }{4} - \theta } \right)} \right],$$
(A.7)

where

$$r_{ij} = \frac{{p_{ij}^{c} }}{\sigma }$$
(A.8)

The perimeter of the non-wetting phase region: \(C_{ij}^{g}\) is given by the following equation:

$$C_{ij}^{g} = 8 \left[ {R_{ij} - r_{ij} \left( { - \frac{{\sin \left( {\frac{\pi }{4} - \theta } \right)}}{{\sin \left( {\frac{\pi }{4}} \right)}} + \left( {\frac{\pi }{4} - \theta } \right)} \right)} \right]$$
(A.9)

The conductance of the wetting liquid phase in the corners of the throat is computed using a semi-empirical model with perfect slip boundary condition between liquid and gas phases (Sidian 2020):

$$K_{ij}^{l} = 4\frac{{2 \tilde{g}_{ij} l_{ij}^{4} }}{{\mu_{ij}^{l} L_{ij} }}$$
(A.10)

\(l_{ij}\), is the meniscus-apex distance along the wall at the corner. It is given by the following equation:

$$l_{ij} = r_{ij} \frac{{\cos \left( {\theta + \beta } \right)}}{\sin \left( \beta \right)}$$
(A.11)

\(\beta\), is half corner angle. For a throat with square cross-sectional area, \(\beta = \frac{\pi }{4}\). The dimensionless conductance of the wetting phase is given by:

$$\tilde{g}_{ij} = {\text{exp}}\left\{ {\frac{{\left[ { - 18.2066 \left( {\tilde{G}_{ij}^{l} } \right)^{2} + 5.88287 \tilde{G}_{ij}^{l} - 0.351809 + 0.02 {\text{sin}}\left( {\beta - \frac{\pi }{6}} \right)} \right]}}{{\left( {\frac{1}{4\pi } - \tilde{G}_{ij}^{l} } \right)}} + 2\ln \tilde{A}_{ij}^{l} } \right\}$$
(A.12)

\(\tilde{A}_{ij}^{l}\) and \(\tilde{G}_{ij}^{l}\) are the area and shape factor of the wetting phase region at a corner with a unit meniscus-apex distance, respectively. \(\tilde{A}_{ij}^{l}\) and \(\tilde{G}_{ij}^{l}\) are given by:

$$\tilde{A}_{ij}^{l} = \left[ {\frac{\sin \left( \beta \right)}{{\cos \left( {\theta + \beta } \right)}}} \right]^{2} \left[ {\frac{{\sin \left( \beta \right)\cos \left( {\theta + \beta } \right)}}{\sin \left( \beta \right)} + \theta + \beta - \frac{\pi }{2}} \right]$$
(A.13)
$$\tilde{G}_{ij}^{l} = \frac{{{ }\tilde{A}_{ij}^{l} { }}}{{4\left[ {1 - \left( {\theta + \beta - \frac{\pi }{2}} \right)\sin \left( \beta \right)/{\text{cos}}\left( {\theta + \beta } \right)} \right]^{2} }}.$$
(A.14)

Appendix B: Algorithm

The algorithm used for solving the molar balance equation and using the fully implicit method is presented in this appendix.

2.1 Algorithm Description

  1. 1.

    Input data and generation of the pore network structure

  2. 2.

    Set initial condition, t = 0, for: \(S^{l}\),\(x^{{{\text{solv}}}}\),\(P^{l}\) in the pores and initial time step, \(\Delta t\).

  3. 3.

    Define the solution vector, the primary variable vector \(X = \left[ {X_{i}^{p} } \right]_{{i = 1,N_{{{\text{pore}}}} }}\), where: For a single liquid-phase pore i, the primary variables are solvent mole fraction and pressure in the liquid phase \(X_{i}^{p} = \left[ {x_{i}^{{{\text{solv}}}} P_{i}^{l} } \right]\). For a two-phase pore i, the primary variables are liquid saturation and pressure: \(X_{i}^{p} = \left[ {S_{i}^{l} P_{i}^{l} } \right]\).

  4. 4.

    While t < tend do

    1. a.

      Initialization of nonlinear iteration loop count: iter = 0

    2. b.

      While \(X^{t + \Delta t,iter + 1} - X^{t + \Delta t,iter}_{2} > tol\) and iter < max_iterations do

    3. c.

      Update throat occupancy based on pore capillary pressures and local rules for throat invasion and imbibition

    4. d.

      Calculate liquid and gas phase conductivities in the throats and their saturation and derivatives with respect to solvent mole fraction and liquid saturation

    5. e.

      Assemble the Jacobian matrix Jac and right-hand side vector F

    6. f.

      Solve the linear system \(Jac \times \Delta X = - F\)

    7. g.

      Update the primary variable vector: \(X^{t + \Delta t,iter + 1} = X^{t + \Delta t,iter} + \Delta X\) and the corresponding liquid-phase saturation, solvent mole fraction and pressure at each pore.

    8. h.

      For a pore under imbibition and liquid saturation sufficiently close to 1.0, make the pore single phase and the throats connected to it all imbibed with the liquid phase

    9. i.

      For a pore under drainage and liquid saturation sufficiently close to \(S_{i min}^{l}\). Fix liquid pore saturation to \(S_{i min}^{l}\) in the time step.

    10. j.

      For pores near phase transition, run flash calculation and determine if the pores are in single or two-phase state. Calculate the saturation of the liquid phase for a two-phase pore.

    11. k.

      Check for convergence of the local newton iteration while checking if the maximum pore saturation change during the iteration exceeds a limiting value,0.2 in this work. If that is the case, reduce the time step by half and restart the iterations at (b).

    12. l.

      When the local newton iterations have converged. Calculate the maximum relative change of pore saturation during the time step. If it exceeds a limiting value, 0.2, reduce the time step by half and restart the iterations at (b).

    13. m.

      Estimate time step \(\Delta t\) for the next time step \(t = t + \Delta t\) using the calculated rate of local liquid drainage or imbibition in the pores. Go back to (a) until t = tend.

Appendix C: Time Step

The time step is estimated using the calculated rate of local liquid drainage or imbibition in the pores and change in solvent mole fraction. Local time step \(\Delta t_{i}\) in pore i is calculated as follows:

  • For two-phase pore i:

    $$\Delta t_{i} = \left\{ {\begin{array}{*{20}l} {\left| {\frac{{\left( {1 - S_{i}^{l} } \right)}}{{q_{i}^{n} }}} \right|, } \hfill & \quad {q_{i}^{n} < 0} \hfill \\ {\left| {\frac{{\left( {S_{i}^{l} - S_{i \min }^{l} } \right)}}{{q_{i}^{n} }}} \right|,} \hfill &\quad {q_{i}^{n} > 0} \hfill \\ \end{array} } \right.$$
    (C.1)

The time step is obtained as follows:

$$\Delta t = \min \left\{ {\left( {\Delta t_{i} } \right)_{{i = 1,N_{{{\text{pore}}}} }} ,2\Delta t_{0} } \right\},$$
(C.2)

where: \(\Delta t_{0}\) is the previous time step.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khammar, M., Niasar, V. Impact of Two-Phase Flow Pattern on Solvent Vapour Extraction. Transp Porous Med 150, 491–517 (2023). https://doi.org/10.1007/s11242-023-01941-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11242-023-01941-5

Keywords

Navigation