Log in

Primary cementing of oil and gas wells in turbulent and mixed regimes

  • Published:
Journal of Engineering Mathematics Aims and scope Submit manuscript

Abstract

We present a detailed derivation of a practical two-dimensional model for turbulent and mixed regimes in narrow annular displacement flows, such as are found during the primary cementing of oil and gas wells. Such mixed cross regimes, including those in which different regimes exist in the same annular cross section, are relatively common in primary cementing. The modelling approach considers scaling based on the disparity of length-scales, which allows a narrow-gap averaging approach to be effective. With respect to the momentum equations, the leading-order equations correspond to a turbulent shear flow in the direction of the modified pressure gradient. This leads to a nonlinear elliptic problem that is the natural extension of the laminar displacement model in Bittleston et al. (J Eng Math 43:229–253, 2002). The mass transport equations that model the miscible displacement are however quite different. To leading-order turbulence effectively mixes the fluids. Changes in concentrations within the annular gap arise due to the combined effects of advection with the mean flow, anisotropic Taylor dispersion (along the streamlines) and turbulent diffusivity. The diffusive and dispersive effects are modelled for fully turbulent and transitional flows following Maleki and Frigaard (J Non-Newt Fluid Mech 235:1–19, 2016). The model derived allows the investigation of different well geometries and inclinations, pum** sequences and fluid rheologies, all of which can have importance. A number of computed examples are presented with the aim of demonstrating the complexity of turbulent displacements.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

Similar content being viewed by others

Notes

  1. A liner is a casing that extends downwards from just above the previous casing.

References

  1. Guillot D, Nelson E (2006) Well Cementing. Schlumberger Educational Services, Sugar Land

    Google Scholar 

  2. Brice JW Jr, Holmes BC (1964) Engineered casing cementing programs using turbulent flow techniques. J Petrol Technol 16:503–508

    Article  Google Scholar 

  3. Couturler M, Guillot D, Hendriks H, Callet F (1990) Design rules and associated spacer properties for optimal mud removal in eccentric annuli. In: CIM/SPE international technical meeting. Society of Petroleum Engineers paper 21594

  4. Jamot A (1974) Déplacement de la boue par le latier de ciment dans l espace annulaire tubage-paroi d un puits. Revue Assoc. Franc. Techn. Petr 224:27–37

    Google Scholar 

  5. Lockyear CF, Ryan DF, Gunningham MM (1990) Cement channeling: how to predict and prevent. Society of Petroleum Engineers, SPE paper 19865

  6. Bittleston SH, Ferguson J, Frigaard IA (2002) Mud removal and cement placement during primary cementing of an oil well- Laminar non-Newtonian displacements in an eccentric annular Hele-Shaw cell. J Eng Math 43:229–253

    Article  MATH  MathSciNet  Google Scholar 

  7. Pelipenko S, Frigaard IA (2004) Two-dimensional computational simulation of eccentric annular cementing displacements. IMA J App Math 69:557–583

  8. Carrasco-Teja M, Frigaard IA, Seymour BR, Storey S (2008) Viscoplastic fluid displacements in horizontal narrow eccentric annuli: stratification and travelling wave solutions. J Fluid Mech 605:293

    Article  ADS  MATH  MathSciNet  Google Scholar 

  9. Pelipenko S, Frigaard IA (2004) On steady state displacements in primary cementing of an oil well. J Eng Math 46:1–26

    Article  MATH  Google Scholar 

  10. Pelipenko S, Frigaard IA (2004) Visco-plastic fluid displacements in near-vertical narrow eccentric annuli: prediction of travelling-wave solutions and interfacial instability. J Fluid Mech 520:343–377

    Article  ADS  MATH  MathSciNet  Google Scholar 

  11. Bogaerts M, Azwar M, Dooply M, Salehpoor A (2015) Well cementing: an integral part of well integrity. In: Offshore Technology Conference Brasil, Rio de Janeiro, Brazil, 27th–29th October 2015. OTC-26314-MS

  12. Gregatti A, Carvalho G, Piedade T, Campos G (2015) Cementing in front of saline zone and carbonate reservoir with presence of \(\text{CO}_\text{2 }\). In: OTC Brasil

  13. Guo Y, Qi J, Zheng YS, Taoutaou S, Wang R, An Y, Guo H (2015) Cementing optimization through an enhanced ultrasonic imaging tool. SPE paper 176069-MS

  14. Osayande V, Isgenderov I, Bogaerts M, Kurawle I, Florez P (2004) Beating the odds with channeling in erd well cementing: solution to persistent channeling problems. Society of Petroleum Engineers, SPE paper 171271-MS

  15. Watson TL (2004) Surface casing vent flow repair-a process. In: Canadian International petroleum conference, Calgary, Alberta, 8th-10th June 2004. PETSOC-2004-297

  16. McLean RH, Manry CW, Whitaker WW (1967) Displacement mechanics in primary cementing. Society of Petroleum Engineers, SPE paper 1488

  17. Allouche M, Frigaard IA, Sona G (2000) Static wall layers in the displacement of two visco-plastic fluids in a plane channel. J. Fluid Mech. 424:243–277

    Article  ADS  MATH  Google Scholar 

  18. Frigaard IA, Leimgruber S, Scherzer O (2003) Variational methods and maximal residual wall layers. J Fluid Mech 483:37–65

    Article  ADS  MATH  MathSciNet  Google Scholar 

  19. Wielage-Burchard K, Frigaard IA (2011) Static wall layers in plane channel displacement flows. J Non-Newt Fluid Mech 166:245–261

    Article  MATH  Google Scholar 

  20. Zare A, Roustaei A, Frigaard IA (2017) Buoyancy effects on micro-annulus formation: Newtonian-bingham fluid displacements in vertical channels. J Non-Newt Fluid Mech (submitted)

  21. Alba K, Taghavi SM, Frigaard IA (2013) Miscible density-unstable displacement flows in inclined tube. Phys Fluids 25:067101

    Article  ADS  Google Scholar 

  22. Taghavi SM, Alba K, Seon T, Wielage-Burchard K, Martinez DM, Frigaard IA (2012) Miscible displacement flows in near-horizontal ducts at low atwood number. J Fluid Mech 696:175–214

    Article  ADS  MATH  MathSciNet  Google Scholar 

  23. Moyers-Gonzalez MA, Frigaard IA (2008) Kinematic instabilities in two-layer eccentric annular flows, part 1: Newtonian fluids. J Eng Math 62:103–131

    Article  MATH  Google Scholar 

  24. Moyers-Gonzalez MA, Frigaard IA (2009) Kinematic instabilities in two-layer eccentric annular flows, part 2: shear-thinning and yield-stress effects. J Eng Math 65:25–52

    Article  MATH  Google Scholar 

  25. Dusseault MB, Jackson RE, Macdonald D (2014) Towards a road map for mitigating the rates and occurrences of long-term wellbore leakage. University of Waterloo and Geofirma Engineering LTD, Waterloo

    Google Scholar 

  26. Watson TL, Bachu S (2008) Identification of wells with high CO\(_2\)-leakage potential in mature oil fields developed for \(\text{ CO }_\text{2 }\)-enhanced oil recovery. In: SPE symposium on improved oil Recovery. Society of Petroleum Engineers

  27. Watson TL, Bachu S (2009) Evaluation of the potential for gas and \(\text{ CO }_\text{2 }\) leakage along wellbores. SPE Drill. Complet. 24:115–126

  28. Carrasco-Teja M, Frigaard IA (2010) Non-Newtonian fluid displacements in horizontal narrow eccentric annuli: effects of slow motion of the inner cylinder. J Fluid Mech 653:137–173

    Article  ADS  MATH  MathSciNet  Google Scholar 

  29. Tardy PMJ, Bittleston SH (2015) A model for annular displacements of wellbore completion fluids involving casing movement. J Pet Sci Eng 126:105–123

    Article  Google Scholar 

  30. Chung SY, Rhee GH, Sung HJ (2002) Direct numerical simulation of turbulent concentric annular pipe flow—Part 1: flow field. Int J Heat Fluid Flow 23:426–440

    Article  Google Scholar 

  31. Freund JB, Lele SK, Moin P (2000) Compressibility effects in a turbulent annular mixing layer. Part 1. Turbulence and growth rate. J Fluid Mech 421:229–267

    Article  ADS  MATH  MathSciNet  Google Scholar 

  32. Nikitin N, Wang H, Chernyshenko S (2009) Turbulent flow and heat transfer in eccentric annulus. J Fluid Mech 638:95–116

    Article  ADS  MATH  Google Scholar 

  33. Nouri JM, Whitelaw JH (1997) Flow of Newtonian and non-Newtonian fluids in an eccentric annulus with rotation of the inner cylinder. Int J Heat Fluid Flow 18(2):236–246

    Article  Google Scholar 

  34. Sawko RAC (2012) Mathematical and computational methods of non-Newtonian, multiphase flows. Ph.D. Thesis, Cranfield University

  35. Maleki A, Frigaard IA (2016) Axial dispersion in weakly turbulent flows of yield stress fluids. J Non-Newt Fluid Mech 235:1–19

    Article  MathSciNet  Google Scholar 

  36. Dodge DW, Metzner AB (1959) Turbulent flow of non-Newtonian systems. AIChE J 5:189–204

    Article  Google Scholar 

  37. Dodge DW, Metzner AB (1962) Errata. AIChE J 8:143

    Google Scholar 

  38. Metzner AB, Reed JC (1955) Flow of non-Newtonian fluids: correlation of the laminar, transition, and turbulent-flow regions. AIChE J 1:434–440

    Article  Google Scholar 

  39. Taghavi SM, Frigaard IA (2013) Estimation of mixing volumes in buoyant miscible displacement flows along near-horizontal pipes. Can J Chem Eng 91:399–412

    Article  Google Scholar 

  40. Zhang J, Frigaard IA (2006) Dispersion effects in the miscible displacement of two fluids in a duct of large aspect ratio. J Fluid Mech 549:225–251

    Article  ADS  MathSciNet  Google Scholar 

  41. Fowler AC (1998) Mathematical models in the applied sciences. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  42. Taylor GI (1954) The dispersion of matter in turbulent flow through a pipe. Proc R Soc Lond A 223:446–468

    Article  ADS  Google Scholar 

  43. Barenblatt GI, Entov VM, Ryzhik VM (1989) Theory of fluid flows through natural rocks. Kluwer Academic, Dordrecht

    MATH  Google Scholar 

  44. Spena FR, Vacca A (2001) A potential formulation of non-linear models of flow through anisotropic porous media. Transp Porous Med 45(3):405–421

    Article  MathSciNet  Google Scholar 

  45. Glowinski R (1984) Numerical methods for nonlinear variational problems. Springer, New York

    Book  MATH  Google Scholar 

Download references

Acknowledgements

This research has been carried out at the University of British Columbia, supported financially by the British Columbia Oil and Gas Commission, by BC OGRIS, and partly by NSERC and Schlumberger through CRD Project 444985-12.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ian Frigaard.

Appendices

Appendix A: The wall shear stress closure

Here we outline the closure relationship that defines the dimensionless \(\tau _w(|\nabla _a \varPsi |;\phi ,\xi ,t)\), locally in the annulus, and hence

$$\begin{aligned} \mathbf {S} = \frac{r_a \tau _w(| \nabla _a \varPsi |) }{H|\nabla _a \varPsi |} \nabla _a \varPsi . \end{aligned}$$

we follow the methods in [35], where the flow of a Herschel–Bulkley fluid along a narrow channel is studied in laminar, transitional, and turbulent regimes.

We start by reconstructing the dimensional variables, which are then scaled following [35]. Given the shear stress scale \(\hat{\tau }_0\), the dimensional wall shear stress is \(\hat{\tau }_w = \hat{\tau }_0 \tau _w\). The dimensional mean speed (averaged across the local gap) and the dimensional local annular gap are given as

$$\begin{aligned} {\hat{W}}_0 = \hat{\bar{W}} \frac{1}{2H} |\nabla _a \varPsi | , ~~~~~~ 2 {\hat{H}} = 2 H \delta _0 {\hat{r}}_{a,0} \end{aligned}$$
(56)

recall that \(\hat{\bar{W}}\) is the velocity scale for the entire annulus. We also assume that, according to the concentrations of the fluids at \((\phi ,\xi ,t)\), we may construct the dimensional local mixture density \(\hat{\rho }\) and the rheological parameters \(\hat{\tau }_{Y}, \hat{\kappa }, n\).

Now, following [35], we define

$$\begin{aligned} \hat{\dot{\gamma }}_{N} = \frac{6 {\hat{W}}_0}{2{\hat{H}}}, ~~~~~~ \hat{\kappa }_p = \hat{\kappa } \left( \frac{2n+1}{3n}\right) ^n. \end{aligned}$$
(57)

The power-law Reynolds number is

$$\begin{aligned} Re_{p} = \frac{6 \hat{\rho } {\hat{W}}_{0}^2}{ \hat{\kappa }_p (\hat{\dot{\gamma }}_{N})^{n}} . \end{aligned}$$
(58)

The Hedström number and scaled wall shear stress are

$$\begin{aligned} He = \hat{\tau }_{Y} \left( \frac{\hat{\rho }^n (2{\hat{H}})^{2n}}{\hat{\kappa }_p^{2}} \right) ^{1/(2-n)}, ~~~~ H_w = \hat{\tau }_{w} \left( \frac{\hat{\rho }^n (2{\hat{H}})^{2n}}{\hat{\kappa }_p^{2}} \right) ^{1/(2-n)} . \end{aligned}$$
(59)

The procedure in [35] gives a detailed description of the map** from \(Re_p\) to \(H_w\) and vice versa, which is parameterized by (nHe). Observe that \(H_w \propto \tau _w\) and \(Re_p \propto |\nabla _a \varPsi |^{2-n}\), so the map** \(H_w \leftrightarrow Re_p\) defines our closure relation. Figure 5 shows an example of this map** for different n at \(He = 100\). The sensitivity to He is not extreme. We now outline the methodology in the different regimes.

1.1 A. 1 Laminar flows

For laminar regime the map** \(H_w \leftrightarrow Re_p\) (i.e. \({\hat{\tau _w}} \leftrightarrow {\hat{W}}_0\)) is

$$\begin{aligned} \frac{(6 Re_{p})^{n/(2-n)}}{H_w} = E\left( n,\frac{He}{H_w}\right) , \end{aligned}$$
(60)

where

$$\begin{aligned} E(n,y_Y) = \left( 1-y_Y\right) ^{(n+1)}\left( \frac{n}{n+1}y_Y+1\right) ^n , \end{aligned}$$
(61)

i.e. \(y_Y = He/H_w = \tau _Y / \tau _w\), which is the dimensionless width of the plug region. This nonlinear relationship is resolved to give \(H_w(Re_p)\) and hence to define \(\mathbf {S}\).

It is noted that \(\mathbf {S}/|\nabla _a \varPsi |\) is singular as \(|\nabla _a \varPsi | \rightarrow 0\), which is the limit where the yield stress of the fluid is not exceeded at the walls of the channel and \(\tau _w < \tau _Y\) is indeterminate. In the laminar displacement model of [6, 7], the vector \(\mathbf {S}\) is defined explicitly to reflect this yielding phenomenon:

$$\begin{aligned}&\mathbf {S} = \left[ \frac{r_a \chi (| \nabla _a \varPsi |) }{|\nabla _a \varPsi |} + \frac{r_a \tau _Y }{H|\nabla _a \varPsi |} \right] \nabla _a \varPsi ~~\Leftrightarrow ~~ |\mathbf {S}| > \frac{r_a \tau _Y }{H}, \end{aligned}$$
(62)
$$\begin{aligned}&| \nabla _a \varPsi | = 0 ~~\Leftrightarrow ~~ |\mathbf {S}| \le \frac{r_a \tau _Y }{H}. \end{aligned}$$
(63)

To connect the model derivation here with that of [6, 7], note that the function \(\chi (| \nabla _a \varPsi |)\) is simply

$$\begin{aligned} \chi (| \nabla _a \varPsi |) = \frac{\tau _w(| \nabla _a \varPsi |) - \tau _Y}{H} , \end{aligned}$$
(64)

where \(\tau _w\) is defined implicitly from (60).

The function \(\chi (| \nabla _a \varPsi |)\) increases strictly monotonically (as does \(\tau _w(| \nabla _a \varPsi |)\)). We can examine the limits of (60), both as \(H_w \rightarrow \infty \) and as \(H_w \rightarrow He\) (yield limit). For the latter, we find

$$\begin{aligned} Re_p^{n/(2-n)} \sim [H_w - He]^{n+1} ~~\Rightarrow ~~ \chi \sim | \nabla _a \varPsi |^{n/(n+1)} ; \end{aligned}$$

see Fig. 5b. As \(H_w \rightarrow \infty \), we find \(Re_p^{n/(2-n)} \sim H_w\), i.e. \(\chi \sim | \nabla _a \varPsi |^n\), reflecting the shear-thinning behaviour. These limiting behaviours agree with those in [7], where the laminar model is analysed in more depth. The difference here though is that the limit \(H_w \rightarrow \infty \) is not physically attained in the laminar regime: we transition to turbulent flow.

1.2 A. 2 Fully turbulent flows

For the fully turbulent regime, in [35] we use the Dodge–Metzner relation, which translates into the following equation defining \(H_w \leftrightarrow Re_p\):

$$\begin{aligned} Re_p = H_w ^{1-\frac{n}{2}} 6^{1-n} 2 ^{1-\frac{n}{2}} \left[ \frac{4.0}{{n'}^{0.75}} \log \left( 6^{1-n'} 2 ^{1-\frac{n'}{2}} E^ {\frac{n'}{n}} H_w^{\frac{n'}{n}- \frac{n'}{2}} \right) -\frac{0.395}{{n'}^{1.2}} \right] ^{2-n} , \end{aligned}$$
(65)

where the generalized power-law index \(n'(n,y_Y)\) is

$$\begin{aligned} n^\prime (n,y_Y) = n(1-y_Y) \frac{ny_Y+n+1}{2n^2y_Y^2+2ny_Y +n+1} \end{aligned}$$
(66)

for \(y_Y= He/H_w\), which represents the dimensionless (effective laminar) plug width. Equation (65) must be solved iteratively for \(H_w\) if \(Re_p\) is specified, but defines \(Re_p\) explicitly if \(H_w\) is specified.

The function \(\tau _w(|\nabla _a \varPsi |)\) is found to increase monotonically in the turbulent regime. Considering He fixed (the rheology) and taking \(H_w \rightarrow \infty \), we find \(y_Y \rightarrow 0, n^{\prime } \rightarrow n\) and \(E \rightarrow 1\). Thus, we find that

$$\begin{aligned} Re_p \sim H_w ^{1-\frac{n}{2}} \log H_w ~~\Rightarrow ~~ | \nabla _a \varPsi | \sim \sqrt{\tau _w} \left[ \log \tau _w \right] ^{1/(2-n)} , \end{aligned}$$

as \(\tau _w \rightarrow \infty \). Thus, \(\tau _w\) grows slightly less fast than \(| \nabla _a \varPsi |^2\), which would be the expectation in a fully rough turbulent regime, and the rheological dependency on n is minimal (in the exponent of the \(\log \) term only), as would also be expected. Thus, we see essentially parallel curves in Fig. 5a at large \(H_w\), independent of n.

1.3 A. 3 Transitional regimes

The transitional regime occurs between two critical values of the Metzner-Reed Reynolds number, \(Re_{MR,1}(n,He/H_w)\) and \(Re_{MR,2}(n,He/H_w)\) [35]. \(Re_{MR,1}\) is the Reynolds number at which the flow is not laminar any more and \(Re_{MR,2}\) is the Reynolds number at which the flow is fully turbulent. Between these values an interpolation is used based on the log of the friction factor. This results in a monotone variation in \(H_w\) vs \(Re_p\), as seen in Fig. 5. \(Re_{MR,1}(n,He/H_w)\) and \(Re_{MR,2}(n,He/H_w)\) can be straightforwardly mapped into \(H_{w,1}\) and \(H_{w,2}\) that bound shear stress for transitional flows.

1.4 A. 4 Friction factor

In passing, we note that in hydraulics, it is common to express closures in terms of a friction factor \(f_f\), defined as the ratio of wall shear stress to inertial stress scale. This can of course be done here, leading to

$$\begin{aligned} \mathbf {S} = \frac{r_a \rho f_f \big | \nabla _a \varPsi \big | }{ 8 H^3} \nabla _a \varPsi . \end{aligned}$$

In [35], the relevant expressions for \(f_f\) are given. For simpler fluids, such as power-law fluids, \(f_f\) is straightforward to evaluate across all regimes, and the Metzner–Reed Reynolds number has proven to be a useful tool. For more complex fluids, the Metzner–Reed Reynolds number is not explicitly defined in terms of the process variables and \(f_f\) is just another level of algebraic complexity that obscures the essential map** between local mean velocity (\(Re_p\)) and wall shear stress (\(H_w\)). The only advantage apparent to us in using \(f_f\) for Herschel–Bulkley fluids is a degree of familiarity with the friction-factor concept.

Fig. 5
figure 5

The closure \(H_w(Re_p)\) from [35], showing asymptotic behaviour: a \(H_w \rightarrow \infty \); b \(H_w \rightarrow He\). The closure is plotted for \(He = 100\) and \(n = 0.2,~0.4,~0.6,~0.8,~1\): green—laminar; red—transitional; black— turbulent. (Color figure online)

Fig. 6
figure 6

Variation of \(\bar{D}_{1D}\)(blue lines), \(D_{T,1D}\)(black lines) and \(D_{T,1D}^*\)(red lines) with wall shear stress for \(n=0.2,0.4,0.6,0.8\) and 1

1.5 A. 5 Evaluating the Taylor dispersion terms

Similar to Sect. 1, we can find the averaged turbulent diffusivity and Taylor dispersion coefficients (\({\bar{D}}, D_T, D_T^*\)) using the method introduced in [35]. We first construct the dimensional parameters, e.g.

$$\begin{aligned} {\hat{D}}_T = D_T \delta _0 {\hat{r}}_a \hat{{\bar{W}}}_0, \end{aligned}$$

and then rescale it using the scaling defined in [35]. We eventually find

$$\begin{aligned} D_T = \frac{1}{2} |\nabla _a \varPsi | D_{T,1D}, \end{aligned}$$
(67)

where \(D_{T,1D}\) is the dispersion coefficient obtained assuming a locally 1D channel flow [35]. Similar relations can be derived for \(\bar{D}\) and \(D_T^*\), i.e. multiplying the 1D results from [35] by \(|\nabla _a \varPsi |/2\).

It is worthwhile to compare the values of \(\bar{D}_{1D},~D_{T,1D}\) and \(D_{T,1D}^*\) (or equivalently \(\bar{D},~D_{T}\) and \(D_{T}^*\)). Figure 6 plots turbulent diffusivity and dispersion coefficients as a function of wall shear stress for fully turbulent flows. \(H_{w,1}\) and \(H_{w,2}\) are defined in Sect.1. As Fig. 6 shows, \({\bar{D}}\) is 2–3 orders of magnitude smaller than \(D_T\). This is a typical feature of turbulent Taylor dispersion [42]. In addition, \(D_T^*\) is almost always larger than \(D_T\). This is interesting, although the results computed so far have not revealed where these terms become important.

Appendix B: Variational principles

Equation (25) is an elliptic second-order equation, in which time evolution enters only via the fluid concentrations (see Sect. 3) or via flow rate changes. Here, we develop the variational theory relevant to solving (25) in a rectangle \(\varOmega \) with boundary \(\partial \varOmega _{\varPsi } \bigcup \partial \varOmega _{S}\), under which conditions (28) and (29), respectively are satisfied. We regard any suitably smooth \(\tilde{\varPsi }\) as an admissible stream function provided that (28) is satisfied. Similarly, \(\tilde{\mathbf {S}}\) will be regarded as admissible provided that

$$\begin{aligned} \mathbf {\nabla }_a \cdot [\tilde{\mathbf {S}} + \mathbf {b}] = 0 , \end{aligned}$$

and that (29) is satisfied. The following statements are easily proven using Green’s theorem in the plane:

  • For any admissible \(\tilde{\varPsi }\) & \(\tilde{\mathbf {S}}\):

    $$\begin{aligned} 0 = \int _{\varOmega } \tilde{\varPsi } \mathbf {\nabla }_a \cdot \mathbf {b} - \nabla _a \tilde{\varPsi } \cdot \tilde{\mathbf {S}}~\mathrm{d}\varOmega + \int _{\partial \varOmega _{\varPsi }} \varPsi _b \tilde{\mathbf {S}}\cdot \mathbf {n}~\mathrm{d}s + \int _{\partial \varOmega _{S}} \tilde{\varPsi } f ~\mathrm{d}s. \end{aligned}$$
    (68)
  • For \(\varPsi \) & \({\mathbf {S}}\) that solve (25) with boundary conditions (28) and (29):

    $$\begin{aligned} 0 = \int _{\varOmega } {\varPsi } \mathbf {\nabla }_a \cdot \mathbf {b} - \nabla _a {\varPsi } \cdot {\mathbf {S}}~\mathrm{d}\varOmega + \int _{\partial \varOmega _{\varPsi }} \varPsi _b {\mathbf {S}}\cdot \mathbf {n}~\mathrm{d}s + \int _{\partial \varOmega _{S}} {\varPsi } f ~\mathrm{d}s . \end{aligned}$$
    (69)
  • For the solution \(\varPsi \) & \({\mathbf {S}}\), and any other admissible \(\tilde{\varPsi }\):

    $$\begin{aligned} 0 = \int _{\varOmega } [\tilde{\varPsi } - \varPsi ] \mathbf {\nabla }_a \cdot \mathbf {b} - [\nabla _a \tilde{\varPsi } - \nabla _a \varPsi ] \cdot \mathbf {S}~\mathrm{d}\varOmega + \int _{\partial \varOmega _{S}} [\tilde{\varPsi }- \varPsi ] f ~\mathrm{d}s . \end{aligned}$$
    (70)
  • For the solution \(\varPsi \) & \({\mathbf {S}}\), and any other admissible \(\tilde{\mathbf {S}}\):

    $$\begin{aligned} \int _{\varOmega } \nabla _a \varPsi \cdot [\tilde{\mathbf {S}} - \mathbf {S}] ~\mathrm{d}\varOmega = \int _{\partial \varOmega _{\varPsi }} \varPsi _b [\tilde{\mathbf {S}}-\mathbf {S}]\cdot \mathbf {n}~\mathrm{d}s . \end{aligned}$$
    (71)

Now we consider the closure relationship defining \(\mathbf {S}\), which is outlined in Appendix A. Provided that \(| \nabla _a \varPsi | > 0\) or equivalently \(|\mathbf {S}| > r_a \tau _Y /H\), we can write this as

$$\begin{aligned} |\mathbf {S}|(| \nabla _a \varPsi |) = \frac{r_a}{H} \tau _w(| \nabla _a \varPsi |) = r_a \chi (| \nabla _a \varPsi |) + \frac{r_a \tau _Y }{H}. \end{aligned}$$
(72)

The function \(\chi (| \nabla _a \varPsi |)\) represents the contribution to the modified pressure gradient that is surplus to that needed to yield the fluid locally. It is continuous and strictly monotone. As the flow transitions through regimes, from laminar through to turbulent, the gradient of \(\chi \) is continuous within any flow regime (but discontinuous when the flow transitions between regimes). Recall that \(\chi (| \nabla _a \varPsi |)\) also has a local dependency on \((\phi ,\xi ,t)\) through the local geometry and fluid concentrations present. However, in general we may represent \(|\mathbf {S}|\) graphically (at any \((\phi ,\xi ,t)\)) as in Fig. 7a.

Fig. 7
figure 7

a \(|\mathbf {S}|(| \nabla _a \varPsi |)\); b \(| \nabla _a \varPsi |(|\mathbf {S}|)\). The shaded areas contribute to the dissipation and potential functions. The shaded areas sum to \(|\mathbf {S}|| \nabla _a \varPsi |\), used to establish Lemma 3

1.1 B. 1 Stream function and pressure potential functionals

The stream function potential functional \(J(\tilde{\varPsi })\) is defined as

$$\begin{aligned} J(\tilde{\varPsi })= & {} \int _{\varOmega } \int _0^{| \nabla _a \tilde{\varPsi } |} |\mathbf {S}|(x) ~\mathrm{d}x - \tilde{\varPsi } \mathbf {\nabla }_a \cdot \mathbf {b} ~\mathrm{d}\varOmega - \int _{\partial \varOmega _{S}} \tilde{\varPsi } f ~\mathrm{d}s , \end{aligned}$$
(73)

which has the following property.

Lemma 1

The solution \(\varPsi \) minimizes \(J(\tilde{\varPsi })\) over all admissible \(\tilde{\varPsi }\).

Proof

We look at:

$$\begin{aligned} J(\tilde{\varPsi })- J({\varPsi })= & {} \int _{\varOmega } \left( \int ^{| \nabla _a \tilde{\varPsi } |}_{| \nabla _a \varPsi |} |\mathbf {S}|(x) ~\mathrm{d}x \right) - (\tilde{\varPsi } - \varPsi ) \mathbf {\nabla }_a \cdot \mathbf {b} ~\mathrm{d}\varOmega - \int _{\partial \varOmega _{S}} (\tilde{\varPsi } - \varPsi ) f ~\mathrm{d}s\\= & {} \int _{\varOmega } \int ^{| \nabla _a \tilde{\varPsi } |}_{| \nabla _a \varPsi |} |\mathbf {S}|(x) ~\mathrm{d}x ~\mathrm{d}\varOmega - \int _{\varOmega } [\nabla _a \tilde{\varPsi } - \nabla _a \varPsi ] \cdot \mathbf {S}(| \nabla _a \varPsi |) ~\mathrm{d}\varOmega \\\ge & {} \int _{\varOmega } \int ^{| \nabla _a \tilde{\varPsi } |}_{| \nabla _a \varPsi |} |\mathbf {S}|(x) ~\mathrm{d}x ~\mathrm{d}\varOmega - \int _{\varOmega } (|\nabla _a \tilde{\varPsi }| - |\nabla _a \varPsi | ) | \mathbf {S}|(| \nabla _a \varPsi |) ~\mathrm{d}\varOmega \\= & {} \int _{\varOmega } \int ^{| \nabla _a \tilde{\varPsi } |}_{| \nabla _a \varPsi |} (|\mathbf {S}|(x) - | \mathbf {S}|(| \nabla _a \varPsi |))~\mathrm{d}x \ge 0 . \end{aligned}$$

We have used (70) and then the Cauchy–Schwarz inequality above. In this last expression, note that \(|\mathbf {S}|(x) > | \mathbf {S}|(| \nabla _a \varPsi |)\) whenever \(x > | \nabla _a \varPsi |\) due to monotonicity. Thus, the sign of the integrant changes according to the limits and the integral is always positive. \(\square \)

The minimization of \(J(\tilde{\varPsi })\) can also be expressed as a variational inequality, which is the basis of the augmented Lagrangian method used. Considering now Fig. 7b, we can define the function \(| \nabla _a \varPsi |(|\mathbf {S}|)\) by effectively inverting \(|\mathbf {S}|(| \nabla _a \varPsi |)\), as illustrated, i.e.

$$\begin{aligned} | \nabla _a \varPsi |(|\mathbf {S}|) = \left\{ \begin{array}{lcl} |\mathbf {S}|^{-1}(| \nabla _a \varPsi |), &{} ~~&{} |\mathbf {S}| > \frac{r_a \tau _Y}{H}, \\ 0, &{} &{} |\mathbf {S}| \le \frac{r_a \tau _Y}{H} . \end{array} \right. \end{aligned}$$

We now define the pressure potential function \(K(\tilde{\mathbf {S}})\) for any admissible \(\tilde{\mathbf {S}}\) as follows.

$$\begin{aligned} K(\tilde{\mathbf {S}}) = - \int _{\varOmega } \int _{\frac{r_a \tau _Y}{H}}^{|\tilde{\mathbf {S}}|} | \nabla _a \varPsi |(y)~\mathrm{d}y ~\mathrm{d}\varOmega + \int _{\partial \varOmega _\varPsi } \varPsi _b \tilde{\mathbf {S}} \cdot \mathbf {n}~\mathrm{d}s. \end{aligned}$$
(74)

Analogous to Lemma 1, we have the following.

Lemma 2

The solution \(\mathbf {S}\) maximizes \(K(\tilde{\mathbf {S}}) \) over all admissible \(\tilde{\mathbf {S}}\).

Proof

We look at

$$\begin{aligned} K(\mathbf {S}) - K(\tilde{\mathbf {S}})= & {} \int _{\varOmega } \int _{|\mathbf {S}|}^{|\tilde{\mathbf {S}}|} | \nabla _a \varPsi |(y)~\mathrm{d}y ~\mathrm{d}\varOmega + \int _{\partial \varOmega _\varPsi } \varPsi _b [\mathbf {S} - \tilde{\mathbf {S}} ] \cdot \mathbf {n}~\mathrm{d}s\\= & {} \int _{\varOmega } \int _{|\mathbf {S}|}^{|\tilde{\mathbf {S}}|} | \nabla _a \varPsi |(y)~\mathrm{d}y ~\mathrm{d}\varOmega - \int _{\varOmega } \nabla _a \varPsi \cdot [\tilde{\mathbf {S}} - \mathbf {S}] ~\mathrm{d}\varOmega \\\ge & {} \int _{\varOmega } \int _{|\mathbf {S}|}^{|\tilde{\mathbf {S}}|} | \nabla _a \varPsi |(y)~\mathrm{d}y ~\mathrm{d}\varOmega - \int _{\varOmega } |\nabla _a \varPsi | (|\tilde{\mathbf {S}}| - |\mathbf {S}|) ~\mathrm{d}\varOmega \\= & {} \int _{\varOmega } \int _{|\mathbf {S}|}^{|\tilde{\mathbf {S}}|} (| \nabla _a \varPsi |(y) - |\nabla _a \varPsi | )~\mathrm{d}y ~\mathrm{d}\varOmega \ge 0 . \end{aligned}$$

Here we have used (71) and then the Cauchy–Schwarz inequality. In the last expression, note that if \(|\tilde{\mathbf {S}}| > | \mathbf {S}|\) then \(| \nabla _a \varPsi |(y) > | \nabla _a \varPsi |\) due to monotonicity; similarly when \(|\tilde{\mathbf {S}}| < | \mathbf {S}|\). Thus, the sign of the integrand changes according to the limits and the integral is always positive. \(\square \)

Finally, since the shaded areas in Fig. 7a, b, sum to give \(|\mathbf {S}| |\nabla _a \varPsi |\), we have

$$\begin{aligned} \int _{\varOmega } |\mathbf {S}| |\nabla _a \varPsi | ~\mathrm{d}\varOmega \int _{\varOmega } \int _0^{| \nabla _a \varPsi |} |\mathbf {S}|(x) ~\mathrm{d}x ~\mathrm{d}\varOmega + \int _{\varOmega } \int _{\frac{r_a \tau _Y}{H}}^{|\mathbf {S}|} | \nabla _a \varPsi |(y)~\mathrm{d}y ~\mathrm{d}\varOmega , \end{aligned}$$

which can be combined with (69). In combination with the above-mentioned minimization and maximization principles, we have the following minimax principle:

Lemma 3

The solution pair \((\mathbf {S}, \varPsi ) \) satisfy

$$\begin{aligned} K(\tilde{\mathbf {S}}) \le K(\mathbf {S}) = J(\varPsi ) \le J(\tilde{\varPsi }) , \end{aligned}$$

for all admissible \(\tilde{\mathbf {S}}\) and \(\tilde{\varPsi }\).

In the porous media context, similar variational principles are used to describe nonlinear filtration, e.g. [43, 44]. The first (integral) terms in both \(J(\cdot )\) and \(K(\cdot )\) are referred to as dissipation potentials. In the porous media context, one is often more concerned with determining the pressure field, and a stream function formulation is restrictive in only applying to 2D flows. Thus, typically \(K(\cdot )\) is referred to as the primal potential and \(J(\cdot )\) as dual potential. Here however, we treat Lemma 1 as the primal principle as it leads to a unique stream function (see below). Note that the terminology dissipation results from (69) which is essentially a mechanical energy balance, equating the dissipation within the system to the work done by buoyancy forces and by the boundary terms.

1.2 B. 2 Existence and uniqueness

Lemma 1 is the basis of an existence and uniqueness result. Firstly, note that \(J(\tilde{\varPsi })\) can be split as follows:

$$\begin{aligned} J(\tilde{\varPsi })= & {} \int _{\varOmega } r_a \int _0^{| \nabla _a \tilde{\varPsi } |} \chi (x) ~\mathrm{d}x ~\mathrm{d}\varOmega + \int _{\varOmega } \frac{r_a \tau _Y}{H} | \nabla _a \tilde{\varPsi } | ~\mathrm{d}\varOmega - \int _{\varOmega } \tilde{\varPsi } \mathbf {\nabla }_a \cdot \mathbf {b} ~\mathrm{d}\varOmega - \int _{\partial \varOmega _{S}} \tilde{\varPsi } f ~\mathrm{d}s , \nonumber \\= & {} J_0(\tilde{\varPsi }) + J_1(\tilde{\varPsi }) - L(\tilde{\varPsi }). \end{aligned}$$
(75)

The functional \(J_0\) is strictly convex as the integrand has second derivative equal to the derivative of \(\chi \), which is a strictly monotone function. The functional \(J_1\) containing the yield stress is convex, bit not strictly. Finally, L denotes the linear parts.

This problem structure is in a format where standard results may be applied (e.g. [45, Theorem 2.1, Chap. 5]), to guarantee the existence of a unique weak solution. The relevant function space is determined by the behaviour of \(J_0\) as \(||\tilde{\varPsi }|| \rightarrow \infty \). From the analysis in Appendix A, we see that in the fully turbulent regime \(| \nabla _a \varPsi | \sim \sqrt{\tau _w} \left[ \log \tau _w \right] ^{1/(2-n)}\) as \(| \nabla _a \varPsi | \rightarrow \infty \) , and therefore also

$$\begin{aligned} \chi \left[ \log \chi \right] ^{2/(2-n)} \sim | \nabla _a \varPsi |^2 . \end{aligned}$$

This suggests \(\chi \gtrsim | \nabla _a \varPsi |^{2 - \epsilon }\) for any small \(\epsilon > 0\) as \(| \nabla _a \varPsi | \rightarrow \infty \), i.e. the \(\log \) term is less significant than any power.

Proceeding now as in [10], we can infer that \(\varPsi \in W^{1,3-\epsilon }(\varOmega )\), with further details specific to the boundary conditions to be considered. It is interesting to compare with the results for the purely laminar case considered in [10], where the growth of \(\chi \) using only the laminar closure resulted in \(\varPsi \in W^{1,1+n_{\min }}(\varOmega )\). It appears that the turbulent closure results in a smoother weak solution and a function space largely independent of the rheology.

1.3 B. 3 Pressure formulation and stress maximization

The maximization of \(K(\tilde{\mathbf {S}})\) leads to an equality as the optimality condition. The resulting partial differential equation (for the pressure) does not, however, uniquely determine \(\mathbf {S}\), where the flow is stationary. We may derive the pressure equation directly by reorganizing (22) to eliminate \(\varPsi \) instead of p:

$$\begin{aligned} 0= & {} \mathbf {\nabla }_a \cdot \left[ r_a^2 \frac{|\nabla _a \varPsi |(|\mathbf {S}|) }{|\mathbf {S}|} \left( \nabla _a p + \mathbf {b}_p \right) \right] , \end{aligned}$$
(76)
$$\begin{aligned} \mathbf {b}_p= & {} \frac{\rho -1}{Fr^2}\left( -\sin \beta \sin \pi \phi , \cos \beta \right) , \end{aligned}$$
(77)
$$\begin{aligned} |\mathbf {S}|= & {} \left| \left( -r_a\frac{\partial \bar{p}}{\partial \xi } - \frac{r_a(\rho -1) \cos \beta }{Fr^2}, \frac{\partial \bar{p}}{\partial \phi } - \frac{r_a(\rho -1) \sin \beta \sin \pi \phi }{Fr^2} \right) \right| . \end{aligned}$$
(78)

The function \(|\nabla _a \varPsi |(|\mathbf {S}|)\) is qualitatively illustrated in Fig. 7b.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Maleki, A., Frigaard, I. Primary cementing of oil and gas wells in turbulent and mixed regimes. J Eng Math 107, 201–230 (2017). https://doi.org/10.1007/s10665-017-9914-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10665-017-9914-x

Keywords

Navigation