Skip to main content
Log in

An Investigation of Polymer Mechanical Degradation in Radial Well Geometry

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

Abstract

It can be challenging to forecast polymer flooding performance at the field, in large part because of the complex non-Newtonian fluid rheology of polymer solutions. In this paper, we apply a model, previously developed to study linear core flooding experiments, to investigate polymer behaviour in radial flow near a vertical injector. The polymer system studied here is very common, HPAM in seawater. One key result is that a grid resolution on the order of millimetres is needed near the wellbore to accurately capture the well pressure, and the amount of mechanical degradation. We also demonstrate that for typical injection rates and permeabilities, the apparent shear thickening and mechanical degradation flow regimes are only relevant to consider within a few metres from the well. For the purposes of full field simulations, a pure shear thinning model should therefore suffice to describe fluid flow outside of the well grid blocks. Approximate analytical expressions are derived to test the numerical model. The steady-state molecular weight far away from the well is shown to scale as \(\propto {Q^{-0.65}\cdot {k^{0.425}}}\), where Q is the injection flow rate, and k is permeability. This scaling makes it possible to collect simulated values onto a single curve and can be used to predict mechanical degradation under different conditions. The results are in broad agreement with observations made of polymer mechanical degradation at the Dalia field. For the case of linear flow, there is an additional length dependency of degradation. The model then predicts an approximate power-law scaling \(M_{\mathrm{w}L}\propto {L^{\omega }}\), with \(M_{\mathrm{w}L}\) being the model molecular weight at a distance L from the inlet, which is consistent with recent laboratory experiments.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

References

  • Åsen, S.M., Stavland, A., Strand, D., Hiorth, A.: An experimental investigation of polymer mechanical degradation at cm and m scale. In: SPE Improved Oil Recovery Conference. Society of Petroleum Engineers (2018)

  • Bird, R.B., Armstrong, R.C., Hassager, O., Curtiss, C.F.: Dynamics of Polymeric Liquids, 1st edn. Wiley, New York (1977)

    Google Scholar 

  • Cannella, W., Huh, C., Seright, R.: Prediction of xanthan rheology in porous media. In: SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers (1988)

  • Carrayrou, J., Mosé, R., Behra, P.: Operator-splitting procedures for reactive transport and comparison of mass balance errors. J. Contam. Hydrol. 68(3–4), 239–268 (2004)

    Article  Google Scholar 

  • Chauveteau, G., Moan, M.: The onset of dilatant behaviour in non-inertial flow of dilute polymer solutions through channels with varying cross-sections. J. Phys. Lett. 42(10), 201–204 (1981)

    Article  Google Scholar 

  • Chauveteau, G., Moan, M., Magueur, A.: Thickening behaviour of dilute polymer solutions in non-inertial elongational flows. J. Non-newtonian Fluid Mech. 16(3), 315–327 (1984)

    Article  Google Scholar 

  • Clemens, T., Lueftenegger, M., Laoroongroj, A., Kadnar, R., Puls, C.: The use of tracer data to determine polymer-flooding effects in a heterogeneous reservoir, 8 Torton Horizon Reservoir, Matzen Field, Austria. SPE Reserv. Eval. Eng. 19(4), 655–663 (2016)

    Article  Google Scholar 

  • Delshad, M., Kim, D.H., Magbagbeola, O.A., Huh, C., Pope, G.A., Tarahhom, F.: Mechanistic interpretation and utilization of viscoelastic behavior of polymer solutions for improved polymer-flood efficiency. In: SPE Symposium on Improved Oil Recovery. Society of Petroleum Engineers (2008)

  • Fletcher, A., Flew, S., Lamb, S., Lund, T., Bjornestad, E., Stavland, A., Gjovikli, N.: Measurements of polysaccharide polymer properties in porous media. In: SPE International Symposium on Oilfield Chemistry. Society of Petroleum Engineers (1991)

  • Ford, J.A.: Improved Algorithms of Illinois-type for the Numerical Solution of Nonlinear Equations. Technical Report CSM-257, University of Essex, Colchester (1995). https://cswww.essex.ac.uk/NA/na_paper.html

  • Glasbergen, G., Wever, D., Keijzer, E., Farajzadeh, R.: Injectivity loss in polymer floods: causes, preventions and mitigations. In: SPE Kuwait Oil and Gas Show and Conference. Society of Petroleum Engineers (2015)

  • Gumpenberger, T., Deckers, M., Kornberger, M., Clemens, T.: Experiments and simulation of the near-wellbore dynamics and displacement efficiencies of polymer injection, Matzen Field, Austria. In: Abu Dhabi International Petroleum Conference and Exhibition. Society of Petroleum Engineers (2012)

  • Herzer, J., Kinzelbach, W.: Coupling of transport and chemical processes in numerical transport models. Geoderma 44(2–3), 115–127 (1989)

    Article  Google Scholar 

  • Howe, A.M., Clarke, A., Giernalczyk, D.: Flow of concentrated viscoelastic polymer solutions in porous media: effect of MW and concentration on elastic turbulence onset in various geometries. Soft Matter 11(32), 6419–6431 (2015)

    Article  Google Scholar 

  • Hundsdorfer, W., Verwer, J.: Numerical solution of advection–diffusion–reaction equations. In: CWI Report NMN9603, Centrum voor Wiskunde en Informatica, Amsterdam, vol. 24, p. 30 (1996)

  • Lantz, R.: Quantitative evaluation of numerical diffusion (truncation error). Soc. Petrol. Eng. J. 11(03), 315–320 (1971)

    Article  Google Scholar 

  • Let, M.S., Priscilla, K., Manichand, R.N., Seright, R.S.: Polymer flooding a \(\sim 500\)-cp oil. In: SPE Improved Oil Recovery Symposium. Society of Petroleum Engineers (2012)

  • Levitt, D.B., Slaughter, W., Pope, G., Jouenne, S.: The effect of redox potential and metal solubility on oxidative polymer degradation. SPE Reserv. Eval. Eng. 14(03), 287–298 (2011)

    Article  Google Scholar 

  • Lohne, A., Nødland, O., Stavland, A., Hiorth, A.: A model for non-Newtonian flow in porous media at different flow regimes. Comput. Geosci. 21(5–6), 1289–1312 (2017)

    Article  Google Scholar 

  • Lueftenegger, M., Kadnar, R., Puls, C., Clemens, T.: Operational Challenges and Monitoring of a Polymer Pilot, Matzen Field, Austria. SPE Production and Operations (2016)

  • Maerker, J.: Shear degradation of partially hydrolyzed polyacrylamide solutions. Soc. Petrol. Eng. J. 15(04), 311–322 (1975)

    Article  Google Scholar 

  • Maerker, J.: Mechanical degradation of partially hydrolyzed polyacrylamide solutions in unconsolidated porous media. Soc. Petrol. Eng. J. 16(04), 172–174 (1976)

    Article  Google Scholar 

  • Manichand, R.N., Let, M.S., Kathleen, P., Gil, L., Quillien, B., Seright, R.S.: Effective propagation of HPAM solutions through the Tambaredjo reservoir during a polymer flood. SPE Prod. Oper. 28(04), 358–368 (2013)

    Google Scholar 

  • Morel, D., Zaugg, E., Jouenne, S., Danquigny, J., Cordelier, P.: Dalia/camelia polymer injection in deep offshore field angola learnings and in situ polymer sampling results. In: SPE Asia Pacific Enhanced Oil Recovery Conference. Society of Petroleum Engineers (2015)

  • Nødland, O.M., Lohne, A., Stavland, A., Hiorth, A.: A model for non-Newtonian flow in porous media at different flow regimes. In: ECMOR XV-15th European Conference on the Mathematics of Oil Recovery (2016)

  • Petzold, L.: Automatic selection of methods for solving stiff and nonstiff systems of ordinary differential equations. SIAM J. Sci. Stat. Comput. 4(1), 136–148 (1983)

    Article  Google Scholar 

  • Ryles, R.: Chemical stability limits of water-soluble polymers used in oil recovery processes. SPE Reser. Eng. 3(01), 23–34 (1988)

    Article  Google Scholar 

  • SciPy.org: scipy.integrate.LSODA. https://docs.scipy.org/doc/scipy/reference/generated/scipy.integrate.LSODA.html. Accessed 04 Sep. 2018

  • Seright, R.S., Seheult, J.M., Talashek, T.: Injectivity characteristics of EOR polymers. In: SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers (2008)

  • Seright, R., Skjevrak, I.: Effect of dissolved iron and oxygen on stability of HPAM polymers. In: SPE Improved Oil Recovery Symposium. Society of Petroleum Engineers (2014)

  • Seright, R.S.: The effects of mechanical degradation and viscoelastic behavior on injectivity of polyacrylamide solutions. Soc. Petrol. Eng. J. 23(03), 475–485 (1983)

    Article  Google Scholar 

  • Seright, R.: Potential for polymer flooding reservoirs with viscous oils. SPE Reserv. Eval. Eng. 13(04), 730–740 (2010)

    Article  Google Scholar 

  • Sharma, A., Delshad, M., Huh, C., Pope, G.A.: A practical method to calculate polymer viscosity accurately in numerical reservoir simulators. In: SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers (2011)

  • Sheng, J.J., Leonhardt, B., Azri, N.: Status of polymer-flooding technology. J. Can. Petrol. Technol. 54(02), 116–126 (2015)

    Article  Google Scholar 

  • Sorbie, K.S.: Polymer-Improved Oil Recovery. Springer, Berlin (1991)

    Book  Google Scholar 

  • Spillette, A., Hillestad, J., Stone, H.: A high-stability sequential solution approach to reservoir simulation. In: Fall Meeting of the Society of Petroleum Engineers of AIME. Society of Petroleum Engineers (1973)

  • Standnes, D.C., Skjevrak, I.: Literature review of implemented polymer field projects. J. Petrol. Sci. Eng. 122, 761–775 (2014)

    Article  Google Scholar 

  • Stavland, A., Jonsbroten, H., Lohne, A., Moen, A., Giske, N.: Polymer flooding-flow properties in porous media versus rheological parameters. In: Presented at the 72nd EAGE Conference and Exhibition Incorporating SPE EUROPEC, SPE-131103-MS (2010)

  • Thomas, A., Gaillard, N., Favero, C.: Some key features to consider when studying acrylamide-based polymers for chemical enhanced oil recovery. Oil Gas Sci. Technol. Revue d’IFP Energies nouvelles 67(6), 887–902 (2012)

    Article  Google Scholar 

  • van den Hoek, P.J., Al-Masfry, R.A., Zwarts, D., Jansen, J.D., Hustedt, B., van Schijndel, L.: Waterflooding under dynamic induced fractures: reservoir management and optimisation of fractured waterfloods. In: SPE Symposium on Improved Oil Recovery. Society of Petroleum Engineers (2008)

  • Wang, D., Han, P., Shao, Z., Hou, W., Seright, R.S.: Sweep-improvement options for the Daqing oil field. SPE Reserv. Eval. Eng. 11(01), 18–26 (2008)

    Article  Google Scholar 

  • Wang, D., Dong, H., Lv, C., Fu, X., Nie, J.: Review of practical experience by polymer flooding at Daqing. SPE Reserv. Eval. Eng. 12(03), 470–476 (2009)

    Article  Google Scholar 

  • Wreath, D., Pope, G., Sepehrnoori, K.: Dependence of polymer apparent viscosity on the permeable media and flow conditions. Situ USA 14(3), 263–284 (1990)

    Google Scholar 

  • Zechner, M., Buchgraber, M., Clemens, T., Gumpenberger, T., Castanier, L.M., Kovscek, A.R.: Flow of polyacrylamide polymers in the near-wellbore-region, rheological behavior within induced fractures and near-wellbore-area. In: SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers (2013)

  • Zechner, M., Clemens, T., Suri, A., Sharma, M.M.: Simulation of polymer injection under fracturing conditions—an injectivity pilot in the Matzen field, Austria. SPE Reserv. Eval. Eng. 18(02), 236–249 (2015)

    Article  Google Scholar 

Download references

Acknowledgements

The authors acknowledge the Research Council of Norway and the industry partners, ConocoPhillips Skandinavia AS, Aker BP ASA, Eni Norge AS, Equinor ASA, Neptune Energy Norge AS, Lundin Norway AS, Halliburton AS, Schlumberger Norge AS, Wintershall Norge AS, and DEA Norge AS, of The National IOR Centre of Norway for support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Oddbjørn Nødland.

Additional information

Publisher's Note

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

Appendices

Appendices

1.1 Derivation of Steady-State Molecular Weight as a Function of Distance

For the molar polymer concentration, a mass balance applied to a control volume V with boundary A yields

$$\begin{aligned} \frac{\partial }{\partial {t}}\int _{V} \phi {C_{\mathrm{mol}}}{{\,\mathrm{d\!}\,}}{V} + \int _{A} C_{\mathrm{mol}}\mathbf {u_p}\cdot \hat{\mathbf {n}}{{\,\mathrm{d\!}\,}}{A} = \int _{V} \phi {{\mathscr {R}}(C_{\mathrm{mol}})}{{\,\mathrm{d\!}\,}}{V}, \end{aligned}$$
(12)

where \(\hat{\mathbf {n}}\) is the outward-pointing unit normal vector to the surface element \({{\,\mathrm{d\!}\,}}{A}\), \(\mathbf {u}\) is the Darcy velocity in vector form, and \(\mathbf {u_p}=\mathbf {u}/(1-\hbox {IPV}_0)\). Dividing by \(\hbox {IPV}_0\) ensures that the correct polymer concentration is transported across the boundary. In differential form, the above equation becomes

$$\begin{aligned} \frac{\partial {(\phi {C_{\mathrm{mol}}})}}{\partial {t}} = -\nabla \cdot ({\mathbf {u}_{p}C_{\mathrm{mol}}}) + \phi {\mathscr {R}}(C_{\mathrm{mol}}). \end{aligned}$$
(13)

The reaction term, given in units of pore volume concentration per time, is

$$\begin{aligned} {\mathscr {R}}(C_{\mathrm{mol}}) = f_{\mathrm{rup}}\cdot {C_{\mathrm{mol}}}. \end{aligned}$$
(14)

At steady-state, the volumetric concentration \(C_{\mathrm{pol}}\) is constant throughout the radial model. For the molar concentration, we obtain

$$\begin{aligned} \nabla \cdot ({\mathbf {u}_{p}C_{\mathrm{mol}}}) = \phi \cdot {f_{\mathrm{rup}}}\cdot {C_{\mathrm{mol}}} \end{aligned}$$
(15)

For radially symmetric flow, we get

$$\begin{aligned} \frac{1}{r}\cdot \frac{\mathrm{d}}{\mathrm{d}r}(ru_{r}C_{\mathrm{mol}})=\phi \cdot {f_{\mathrm{rup}}} \cdot {C_{\mathrm{mol}}} \end{aligned}$$
(16)

Inserting \(u_r=Q/(2\pi {r}h(1-\hbox {IPV}_0)\) yields

$$\begin{aligned} \frac{Q}{2\pi {h}r(1-\hbox {IPV}_0)}\cdot \frac{\mathrm{d}C_{mol}}{\mathrm{d}r} =\phi \cdot {f_{\mathrm{rup}}}\cdot {C_{\mathrm{mol}}}, \end{aligned}$$
(17)

and using that \(M_\mathrm{w}=C_{\mathrm{pol}}/C_{\mathrm{mol}}\) we obtain

$$\begin{aligned} \frac{\mathrm{d}C_{\mathrm{mol}}}{\mathrm{d}r}=-\frac{C_{\mathrm{pol}}}{M_{\mathrm{w}}^2}\cdot \frac{\mathrm{d}M_\mathrm{w}}{\mathrm{d}r} =\frac{2\pi {h}r\phi (1-\hbox {IPV}_0)}{Q}\cdot {f_{\mathrm{rup}}}\cdot {C_{\mathrm{mol}}}, \end{aligned}$$
(18)

from which Eq. (8) follows. For the case of linear 1D geometry, we can repeat the above procedure to obtain an identical kind of formula,

$$\begin{aligned} \frac{\mathrm{d}M_\mathrm{w}}{\mathrm{d}x} = -\frac{A_{l}\phi (1-\hbox {IPV}_0)}{Q}\cdot {f_{\mathrm{rup}}}\cdot {M_\mathrm{w}}, \end{aligned}$$
(19)

where \(A_l\) is the constant, cross-sectional area of the core.

1.2 Approximate Analytical Formulas for Degraded \(M_\mathrm{w}\)

Consider radial flow outwards from an injector at flow rate Q, in a homogeneous reservoir. Based on the derivations in the previous section, we see that at steady state, the molecular weight as a function of radial distance r must satisfy Eq. (8) in the main text, where \(f_{\mathrm{rup}}\) is given by Eq. (6). The formula used for apparent viscosity in the simulation model is

$$\begin{aligned} \eta&=\eta _s + (\eta _{\mathrm{sh}}-\eta _s)\cdot {\eta _{\mathrm{elf}}} \nonumber \\&\approx {(\eta _{\mathrm{sh}}-\eta _s)\cdot {\eta _{\mathrm{elf}}}} \nonumber \\&=\eta _{s}\cdot \eta _{\mathrm{sp},\mathrm{sh}}\cdot \eta _{\mathrm{elf}}\nonumber \\&=\eta _{s}\cdot {\eta _{\mathrm{sp}0}}\cdot (1+(\lambda _{1}{\dot{\gamma }})^{x})^{-\frac{n}{x}} \cdot (1+(\lambda _{2}{\dot{\gamma }})^{x_2})^{\frac{m+n}{x_2}}, \end{aligned}$$
(20)

where \(\eta _{\mathrm{sp},\mathrm{sh}}=\frac{\eta _{\mathrm{sh}}}{\eta _s}-1\) is the specific viscosity for the shear thinning part of the apparent viscosity, and \(\eta _{\mathrm{elf}}\) is the elongational viscosity factor. In the second line, we have made an approximation by dropping the solvent viscosity term in the expression for \(\eta \). The zero shear specific viscosity, \(\eta _{\mathrm{sp}0}\), is calculated from a cubic polynomial in the product of intrinsic viscosity and polymer concentration,

$$\begin{aligned} \eta _{\mathrm{sp}0} = [\eta ]C_{\mathrm{p}}+k^{\prime }[\eta ]^{2}C_{\mathrm{p}}^2 + k^{\prime \prime }[\eta ]^{3}C_{\mathrm{p}}^3, \end{aligned}$$
(21)

with the intrinsic viscosity obtained from the molecular weight via the Mark–Houwink equation,

$$\begin{aligned}{}[\eta ]=K\cdot {M_{\mathrm{w}}^a}, \end{aligned}$$
(22)

for constants K and a. In terms of the introduced notation and the approximation introduced above, the viscosity term in (7) becomes

$$\begin{aligned} \eta ^{\alpha _d}\approx {\eta _{s}^{\alpha _d}}\cdot {\eta _{\mathrm{sp}0}}^{\alpha _d} \cdot (1+(\lambda _{1}{\dot{\gamma }})^{x})^{-\frac{n\alpha _d}{x}} \cdot (1+(\lambda _{2}{\dot{\gamma }})^{x_2})^{\frac{(m+n)\alpha _d}{x_2}}. \end{aligned}$$
(23)

Next, since in the degradation regime \({\dot{\gamma }}\gg {1}\), we assume that

$$\begin{aligned} (1+(\lambda _{1}{\dot{\gamma }})^{x})^{-\frac{n}{x}} \approx (\lambda _{1}{\dot{\gamma }})^{-n}, \end{aligned}$$
(24)

and

$$\begin{aligned} (1+(\lambda _{2}{\dot{\gamma }})^{x_2})^{-\frac{m+n}{x_2}} \approx (\lambda _{2}{\dot{\gamma }})^{m+n}. \end{aligned}$$
(25)

The expressions for n, \(\lambda _1\), and \(\lambda _2\) are given by, respectively, Eqs. 4, 21, and 30 in Lohne et al. (2017). We reproduce those equations here:

$$\begin{aligned} n=n(M_\mathrm{w})=1-\frac{1}{1+\left( a_{n}\hbox {KC}_{\mathrm{p}}M_{\mathrm{w}}^a\right) ^{b_n}}, \end{aligned}$$
(26)

and

$$\begin{aligned} \lambda _{1}=\lambda _{1}(M_\mathrm{w})=\lambda _{a}\cdot \frac{\eta _{s}\eta _{\mathrm{sp}0}(M_\mathrm{w})M_\mathrm{w}}{C_{\mathrm{p}}T}, \end{aligned}$$
(27)

and

$$\begin{aligned} \lambda _{2}=\lambda _{2}(M_\mathrm{w})=\frac{1}{N_{\mathrm{De}}^{\star }}\cdot \frac{3}{5R_g} \cdot \frac{\phi }{1-\phi }\cdot \frac{\eta _{s}KM_{\mathrm{w}}^{a+1}}{T}. \end{aligned}$$
(28)

In the degradation regime, the main contribution to the viscosity is from the shear thickening part. Hence, to make the analysis tractable we assume in the sequel that n is constant, i.e., independent of \(M_\mathrm{w}\). With all these approximations, Eq. (23) reduces to

$$\begin{aligned} \eta ^{\alpha _d}&\approx {\eta _{s}^{\alpha _d}}\cdot {\eta _{\mathrm{sp}0}}^{\alpha _d} \cdot \lambda _{1}^{-n\alpha _d}\cdot \lambda _{2}^{(m+n)\alpha _d}\cdot {{\dot{\gamma }}^{m\alpha _d}} \nonumber \\&=\eta _{s}^{\alpha _d}\cdot {\eta _{\mathrm{sp}0}^{\alpha _{d}(1-n)}}\cdot \left( \frac{\lambda _{a}\eta _{s}M_\mathrm{w}}{C_{\mathrm{p}}T}\right) ^{-n\alpha _d} \cdot \left( \frac{3}{5N_{\mathrm{De}}^{\star }R_g}\cdot \frac{\phi }{1-\phi }\cdot \frac{\eta _{s}KM_{\mathrm{w}}^{1+a}}{T}\right) ^{(m+n)\alpha _d} \cdot {{\dot{\gamma }}^{m\alpha _d}}, \end{aligned}$$
(29)

Furthermore, we only include the first term in the expression for \(\eta _{\mathrm{sp}0}\),

$$\begin{aligned} \eta _{\mathrm{sp}0}\approx {f_{1}(M_\mathrm{w})}\equiv {\hbox {KC}_{\mathrm{p}}M_{\mathrm{w}}^a}. \end{aligned}$$
(30)

The last two approximations are the boldest, especially Eq. (30), however as we shall see, without them it is not possible to integrate the degradation equation in terms of elementary functions. By combining Eqs. (3), (7), (8), (29) and (30), we obtain

$$\begin{aligned} \frac{\mathrm{d}M_\mathrm{w}}{\mathrm{d}r}&\approx {-\frac{2\pi {h}r\phi (1-\hbox {IPV}_0)}{Q}} \cdot \left( r_{\mathrm{deg}}\eta _{s}K{C_\mathrm{p}}\right) ^{\alpha _d} \cdot \sqrt{\frac{\phi }{2kC}}\nonumber \\&\quad \cdot \sqrt{R_{k}(1-\hbox {IPV}_0)} \cdot \left( \frac{\lambda _{a}\eta _{s}K}{T}\right) ^{-n\alpha _d}\nonumber \\&\quad \cdot \left( \frac{3}{5N_{\mathrm{De}}^{\star }R_g}\cdot \frac{\phi }{1-\phi }\cdot \frac{\eta _{s}K}{T}\right) ^{(m+n)\alpha _d} \cdot {M_{\mathrm{w}}^y}\cdot \left( \frac{\Omega }{r}\right) ^{\alpha _{d}(1+m)}, \end{aligned}$$
(31)

with y given by

$$\begin{aligned} y=1+\beta _d+\alpha _{d}(a(1+m)+m), \end{aligned}$$
(32)

and where we have defined \(\Omega \) as the part of \({\dot{\gamma }}\) that does not depend on r. From Eq. (2), using that \(u=Q/2\pi {r}h\), this means that

$$\begin{aligned} \Omega = \frac{4\alpha _{c}Q}{2\pi {h}\sqrt{8k{\phi }}}\cdot {\sqrt{\frac{R_\mathrm{k}}{1-\hbox {IPV}_0}}}. \end{aligned}$$
(33)

By collecting all terms other than \(M_\mathrm{w}\) and r into a factor \(\zeta \), it is seen that we approximate the original problem by the separable ODE

$$\begin{aligned} \frac{\mathrm{d}M_\mathrm{w}}{\mathrm{d}r}={-\zeta }\cdot {M_{\mathrm{w}}^y}\cdot {r^{\mathrm{w}}}, \end{aligned}$$
(34)

with \(w=1-\alpha _{d}(1+m)\). Let \(r_d\) be the radius beyond which there is no more, or negligible, mechanical degradation. By substituting \({\dot{\gamma }}=\Omega /r\), and integrating from \(r_\mathrm{w}\) to \(r_d\), we get

$$\begin{aligned} \int _{M_{\mathrm{w}0}}^{M_{\mathrm{wd}}} {M_{\mathrm{w}}^{-y}}{{\,\mathrm{d\!}\,}}{M_\mathrm{w}}=\Omega ^{1+w}\zeta \cdot \int _{{\dot{\gamma }}_{w}}^{{\dot{\gamma }}_d} {\frac{1}{{\dot{\gamma }}^{2+w}}}{{\,\mathrm{d\!}\,}}{{\dot{\gamma }}}, \end{aligned}$$
(35)

where \(M_{\mathrm{wd}}=M_{\mathrm{w}}(r_d)\) is the steady-state molecular weight far away from the injection well, and \(M_{\mathrm{w}0}\) is the initial molecular weight. From this, it immediately follows that

$$\begin{aligned} \frac{1}{1-y}\cdot \left( M_{\mathrm{wd}}^{1-y}-M_{\mathrm{w}0}^{1-y}\right) = \frac{-\Omega ^{1+w}\zeta }{1+w}\cdot \left( {\dot{\gamma }}_{d}^{-(1+w)}-{\dot{\gamma }}_{w}^{-(1+w)}\right) . \end{aligned}$$
(36)

By virtue of the definition of \(r_d\), the shear rate \({\dot{\gamma }}_d\) will be negligible compared with \({\dot{\gamma }}_w\). Thus, as a final approximation, we will assume\({\dot{\gamma }}_{d}^{-(1+w)}\approx {0}\). This is justified by comparing with the actual simulation results in the radial grid. Finally, by performing the necessary algebraic manipulations, one can show that

$$\begin{aligned} \frac{M_{\mathrm{wd}}}{M_{\mathrm{w}0}}\approx \frac{1}{(1+(y-1)\chi _r)^{\frac{1}{y-1}}}, \end{aligned}$$
(37)

where \(\chi _r\) is the following complicated expression:

$$\begin{aligned} \chi _r = \chi _{0}\cdot {(1-\hbox {IPV}_0)\cdot {R_\mathrm{k}}}\cdot \frac{\phi ^{1+(m+n)\alpha _d}}{(1-\phi )^{(m+n)\alpha _d}} \cdot {C_{\mathrm{p}}^{\alpha _d}}\cdot {T^{-m\alpha _d}}\cdot {r_\mathrm{w}}\cdot {M_{\mathrm{w}0}^{y-1}} \cdot \frac{{\dot{\gamma }}_{w}^{\alpha _{d}(1+m)-1}}{k},\nonumber \\ \end{aligned}$$
(38)

with

$$\begin{aligned} \chi _0 =\frac{\alpha _c}{\sqrt{C}}\cdot \frac{1}{\alpha _{d}(1+m)-2} \cdot {r_{\mathrm{deg}}}^{\alpha _d} \cdot {\lambda _{a}}^{-n\alpha _d} \cdot \left( \frac{3}{5N_{\mathrm{De}}^{\star }{R_g}}\right) ^{(m+n)\alpha _d} \cdot (\eta _{s}K)^{\alpha _{d}(1+m)}.\qquad \end{aligned}$$
(39)
Table 2 Rock and fluid properties used as input parameters for the radial simulations

For the particular choice of model parameters used in this paper (Tables 2 and 3), we find that \(\chi _{r}\propto {\frac{{\dot{\gamma }}_{w}^{6.5}}{k}}\). By inserting the definition of \({\dot{\gamma }}_w\), and collecting equal terms, one can further reduce (38) to

$$\begin{aligned} \chi _r= & {} \chi _{0}\cdot {\left( \frac{\alpha _c}{\sqrt{2}\pi }\right) ^{\alpha _{d}(1+m)-1}} \cdot {Q^{a_1}}\cdot {k^{a_2}}\cdot {T^{a_3}}\cdot {\phi ^{a_4}}\cdot {(1-\phi )^{a_5}}\nonumber \\&\cdot ~{(1-\hbox {IPV}_0)^{a_6}}\cdot {R_{k}^{a_7}}\cdot {r_{w}^{a_8}}\cdot {h^{a_9}} \cdot {M_{\mathrm{w}0}^{a_{10}}} \cdot {C_{\mathrm{p}}^{a_{11}}}, \end{aligned}$$
(40)

with the exponents given in Table 4. We remark that the temperature dependence is not fully captured by the T-term in Eq. (40), as we also have \(\eta _s=\eta _{s}(T)\) and \(K=K(T)\) in the term \(\chi _0\). For very large \(\chi _r\), Eq. (37) becomes

$$\begin{aligned} \frac{M_{\mathrm{wd}}}{M_{\mathrm{w}0}}\approx ((y-1)\cdot \chi _{r})^{-\frac{1}{y-1}} \end{aligned}$$
(41)
Table 3 Polymer properties used as input to the simulations
Table 4 Exponents appearing in Eq. (40)

1.3 Alternative Approximate Equations

Returning to Eq. (29), let us approximate \(\eta _{sp0}\) in a different way than in the development of Eq. (37). Let R denote the ratio between the alternative formula and the original one, i.e.,

$$\begin{aligned} R\equiv \frac{\eta _{\mathrm{sp}0}}{f_{1}}=\frac{\eta _{\mathrm{sp}0}}{\hbox {KC}_{\mathrm{p}}M_{\mathrm{w}}^a}. \end{aligned}$$
(42)

Then, from Eq. (29) it is clear that the right-hand side of Eq. (34) must be multiplied by a factor \(R^{\alpha _{d}(1-n)}\). Equivalently, when integrating the \(M_\mathrm{w}\) and r terms, the integrand on the left-hand side of Eq. (35) must be multiplied by a factor \(R^{\alpha _{d}(n-1)}\).

1.3.1 Power-Law

If the factor R is proportional to a power of \(M_{\mathrm{w}}\), the integration can be performed in exactly the same way as before, and we end up with the same kind of formula as (37). The only difference is that the definition of \(\chi _r\) must be modified with an extra prefactor, in addition to changing the exponent y. For instance, we can assume that \(\eta _{\mathrm{sp}0}\approx {f_{3}}\), where \(f_3\) is the third-order term in Eq. (21). An example of using the latter assumption is shown in the left plot of Fig. 4 (green curve).

1.3.2 Including Both the First and the Third-Order Term in \(\eta _{\mathrm{sp}0}\)

Another possibility is to improve the approximation of the cubic formula by only disregarding the quadratic term (which makes the smallest contribution). In this case, we get

$$\begin{aligned} R^{\alpha _{d}(1-n)}=\frac{f_1+f_3}{f_1}= 1+\frac{f_3}{f_1}=1+k^{\prime \prime }K^{2}C_{\mathrm{p}}^{2}M_{\mathrm{w}}^{2a}, \end{aligned}$$

resulting in the approximate equation

$$\begin{aligned} \int _{M_{\mathrm{w}0}}^{M_{\mathrm{wd}}}{(1+BM_{\mathrm{w}}^{2a})^{\alpha _{d}(n-1)} \cdot {}M_{\mathrm{w}}^{-y}}{{\,\mathrm{d\!}\,}}{M_\mathrm{w}}=\Omega ^{1+w}\zeta \cdot \int _{{\dot{\gamma }}_{w}}^{{\dot{\gamma }}_d} {\frac{1}{{\dot{\gamma }}^{2+w}}}{{\,\mathrm{d\!}\,}}{{\dot{\gamma }}}, \end{aligned}$$
(43)

with \(B=k^{\prime \prime }K^{2}C_{\mathrm{p}}^{2}\). Substituting \(u=B\cdot {M_{\mathrm{w}}^{2a}}\), we transform the integral on the left-hand side to

$$\begin{aligned} \frac{B^{\frac{y-1}{2a}}}{2a} \int _{u_0}^{u_d} (1+u)^{C}\cdot {u^D}{{\,\mathrm{d\!}\,}}{u} \end{aligned}$$

with \(C=(n-1)\alpha _d\), \(D=(1-y-2a)/2a\), \(u_0=B\cdot {M_{\mathrm{w}0}^{2a}}\), and \(u_d=B\cdot {M_{\mathrm{w}d}^{2a}}\). This definite integral may be expressed in terms of the Gaussian hypergeometric function \(_{2}F_{1}\),

The left-hand side of Eq. (43) now becomes

If we denote the hypergeometric function evaluated at the upper and lower limits by, respectively, \({\mathscr {F}}_d\) and \({\mathscr {F}}_0\), we obtain the following approximate relationship:

$$\begin{aligned} M_{\mathrm{wd}}^{1-y}\cdot {{\mathscr {F}}_d}-M_{\mathrm{w}0}^{1-y}\cdot {{\mathscr {F}}_0}\approx \frac{(1-y)\Omega ^{1+w}\zeta }{1+w}\cdot {\dot{\gamma }}_{w}^{-(1+w)}. \end{aligned}$$
(44)

To isolate \(M_{\mathrm{wd}}\) we need to compute the inverse of \(_{2}F_{1}\), which we do by numerically solving the implicit Eq. (44).

1.4 Analytical Solution by Means of Numerical Integration

As documented in the main text, the use of Eq. (37) underestimates the amount of degradation. This is in large part due to neglecting the third-order term in Eq. (21). The assumption of a constant Carreau–Yasuda exponent n also explains a part of the discrepancy between simulation and calculation. This can be verified by performing numerical integration of Eq. (8). Tracing through the various definitions, one can show that the integral we need to solve is

$$\begin{aligned} \frac{\mathrm{d}M_\mathrm{w}}{\mathrm{d}r} = -\frac{2{\pi }h\phi (1-\hbox {IPV}_0)\Omega ^{\alpha _d}}{Q\sqrt{2C}}\cdot \sqrt{\frac{\phi }{k}} \cdot \sqrt{R_{k}(1-\hbox {IPV}_0)}\cdot (r_{\mathrm{deg}}\eta _s)^{\alpha _d}\cdot {r^{1-\alpha _d}}\cdot {\eta _{r}^{\alpha _d}} \cdot {M_{\mathrm{w}}^{1+\beta _d}}, \nonumber \\ \end{aligned}$$
(45)

where the relative apparent viscosity \(\eta _r\) is given by

$$\begin{aligned} \eta _r = 1 + \eta _{\mathrm{sp}0}\cdot {(1+(\lambda _{1}{\dot{\gamma }})^{x})^{-n/x}}\cdot (1+(\lambda _{2}{\dot{\gamma }})^{x_2})^{(m+n)/x_2}. \end{aligned}$$
(46)

Numerical integration was performed using Python’s odeint function, which is essentially a wrapper to the LSODA solver in the FORTRAN library odepack. According to the online documentation (SciPy.org 2018), the integrator switches automatically between non-stiff (Adams) and stiff (backwards differentiation formulas, BDF) methods, depending on information available at the end of each integration step. See, e.g., Petzold (1983) for more details regarding this method.

1.5 Linear Geometry and Approximate Scaling Relationships

The procedure used to derive Eqs. (37) and (45) may also be applied to the case of linear 1D geometry by using Eq. (19). Let \(M_{\mathrm{w}L}\) denote the steady-state molecular weight at \(x=L\). By making the same assumptions as before, one can show that

$$\begin{aligned} \frac{M_{\mathrm{w}L}}{M_{\mathrm{w}0}}\approx \frac{1}{(1+(y-1)\chi _l)^{\frac{1}{y-1}}}, \end{aligned}$$
(47)

where \(\chi _{_{l}}\) is similar to \(\chi _r\) (not shown here). As a further test of the derived analytical expressions, we conducted a new series of simulations, this time in linear geometry at the core scale. For each simulation, the core permeability was chosen to be the same as in a corresponding radial case, and the flow rate was selected so that \({\dot{\gamma }}_{_{l}}={\dot{\gamma }}_w\), where \({\dot{\gamma }}_{_{l}}\) is the in situ shear rate in the core. The \(L=7~\hbox {cm}\) core was discretized into 100 grid blocks, i.e., with a constant grid spacing \(\Delta {x}=0.7~\hbox {mm}\). Figure 12 shows a comparison between the numerical integration results, and values obtained from the explicit first-order formula and the simulations. As discussed in the main text, the approximate formulas tend to overpredict \(M_\mathrm{w}\). For the core scale simulations, very good agreement between simulation and numerical integration was obtained, the maximal relative error being less than 1%.

Fig. 12
figure 12

Left plot: Simulated and calculated molecular weight at the exterior grid boundary, plotted versus the exact solution (numerical integration, Eq. (45)). The red triangles are obtained from the analytical formula, Eq. (37), while the blue circles represent the simulations. Right plot: Simulated and calculated molecular weight at the core effluent for a series of linear simulations, plotted versus the exact solutions

Note that under the assumptions used to derive Eq. (47), the model predicts an approximate power-law scaling for \(M_\mathrm{w}\) as a function of distance in linear flow, all else being equal. Specifically, if the constant 1 can be neglected in the denominator, we have

$$\begin{aligned} \frac{M_{\mathrm{w}L}}{M_{\mathrm{w}0}}\propto {L^\omega }, \end{aligned}$$
(48)

with \(\omega =1-y=-0.1\) in the present case. If approximating \(\eta _{\mathrm{sp}0}\) with the third-order term rather than the first-order term (Sect. C.1), a slightly different exponent is obtained, but power-law scaling was seen for all considered cases in which degradation was substantial.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nødland, O., Lohne, A., Stavland, A. et al. An Investigation of Polymer Mechanical Degradation in Radial Well Geometry. Transp Porous Med 128, 1–27 (2019). https://doi.org/10.1007/s11242-018-01230-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11242-018-01230-6

Keywords

Navigation