Skip to main content
Log in

A hybrid asymptotic-numerical method for calculating drag coefficients in 2-D low Reynolds number flows

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

Abstract

Steady-state incompressible low-Reynolds-number fluid flow past a cylindrical body in an unbounded two-dimensional domain is a singular perturbation problem involving an infinite logarithmic expansion in the Reynolds number \({\displaystyle \varepsilon }\) as \({\displaystyle \varepsilon }\rightarrow 0\). The central difficulty with applying a conventional matched asymptotic approach to this problem is that only the first few terms in the infinite logarithmic expansion of the drag coefficient and of the flow field can be calculated analytically. To overcome this difficulty, a hybrid asymptotic-numerical method that incorporates all logarithmic correction terms is implemented for three low-Reynolds-number flow problems. In particular, for a nanocylinder of circular cross section with surface roughness, modeled by a Navier boundary condition involving a sliplength parameter, a hybrid asymptotic-numerical method is formulated and implemented to determine an approximation to the drag coefficient that is accurate to all powers of \({-1/\log {\displaystyle \varepsilon }}\). A similar analysis is done to determine a corresponding approximation of the drag coefficient for a porous cylinder, where the flow inside the cylinder is modeled by the Brinkman equation. For both the nano- and porous-cylinder problems, the hybrid asymptotic-numerical method is extended to calculate the first transcendentally small correction term to the Stokes flow near a body. This term, which governs weak upstream/downstream asymmetry in the Stokes flow, is extrapolated to finite \({\displaystyle \varepsilon }\) to predict the formation of any eddies near the body. Finally, the hybrid method is used to determine the drag coefficient, valid to within all logarithmic terms, for two identical cylinders of circular cross section in tandem alignment with the free stream. An extension of the theoretical framework to more general slow viscous flow problems is discussed.

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

References

  1. Kaplun S (1957) Low Reynolds number flow past a circular cylinder. J Math Mech 6(5):52–60

    MathSciNet  MATH  Google Scholar 

  2. Proudman I, Pearson J (1957) Expansions at small Reynolds number for the flow past a sphere and a circular cylinder. J Fluid Mech 2:237–262

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Tritton DJ (1959) Experiments on the flow past a circular cylinder at low Reynolds numbers. J Fluid Mech 6:547–567

    Article  ADS  MATH  Google Scholar 

  4. Van Dyke M (1975) Perturbation methods in fluid mechanics. Parabolic Press, Stanford

    MATH  Google Scholar 

  5. O’Malley RE (2010) Singular perturbation theory: a viscous flow out of Göttingen. Annu Rev Fluid Mech 42:1–17

    Article  ADS  MATH  Google Scholar 

  6. Veysey J II, Goldenfeld N (2007) Simple viscous flows: from boundary layers to the renormalization group. Rev Mod Phys 79:883–927

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Kropinski MC, Ward MJ, Keller JB (1995) A hybrid asymptotic-numerical method for low Reynolds number flows past a cylindrical body. SIAM J Appl Math 55(6):1484–1510

    Article  MathSciNet  MATH  Google Scholar 

  8. Lee SH, Leal LG (1986) Low Reynolds number flow past cylindrical bodies of arbitrary cross-sectional shape. J Fluid Mech 164:401–427

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Titcombe M, Ward MJ, Kropinski MC (2000) A hybrid asymptotic-numerical solution for low Reynolds number flow past an asymmetric cylindrical body. Stud Appl Math 105(2):165–190

    Article  MathSciNet  MATH  Google Scholar 

  10. Ward MJ, Henshaw WD, Keller JB (1993) Summing logarithmic expansions for singularly perturbed eigenvalue problems. SIAM J Appl Math 53(3):799–828

    Article  MathSciNet  MATH  Google Scholar 

  11. Kolokolnikov T, Titcombe M, Ward MJ (2005) Optimizing the fundamental Neumann eigenvalue for the laplacian in a domain with small traps. Eur J Appl Math 16(2):161–200

    Article  MathSciNet  MATH  Google Scholar 

  12. Pillay S, Ward MJ, Pierce A, Kolokolnikov T (2010) An asymptotic analysis of the mean first passage time for narrow escape problems: part I: two-dimensional domains. SIAM Multiscale Mod Sim 8(3):803–835

    Article  MATH  Google Scholar 

  13. Kropinski MC, Lindsay A, Ward MJ (2011) Asymptotic analysis of localized solutions to some linear and nonlinear biharmonic eigenvalue problems. Stud Appl Math 126(4):347–408

    Article  MathSciNet  MATH  Google Scholar 

  14. Chen W, Ward MJ (2011) The stability and dynamics of localized spot patterns in the two-dimensional gray-scott model. SIAM J Appl Dyn Syst 10(2):582–666

    Article  MathSciNet  MATH  Google Scholar 

  15. Ward MJ, Kropinski MC (2010) Asymptotic methods for PDE problems in fluid mechanics and related systems with strong localized perturbations in two dimensional domains. In: Steinrück H (ed) Asymptotic methods in fluid mechanics: surveys and results. CISM International Centre for Mechanical Science, vol 523. Springer, New York, pp 23–70

  16. Matthews MT, Hill JM (2006) Flow around nanospheres and nanocylinders. Q J Mech Appl Math 59(2):191–210

    Article  MathSciNet  MATH  Google Scholar 

  17. Kohr M, Prakash J, Sekhar GPR, Wendland WL (2009) Expansions at small Reynolds numbers for the flow past a porous circular cylinder. Appl Anal 88(7):1093–1114

    Article  MathSciNet  MATH  Google Scholar 

  18. Keller JB, Ward MJ (1996) Asymptotics beyond all orders for a low Reynolds number flow. J Eng Math 30:253–265

    Article  MathSciNet  MATH  Google Scholar 

  19. Hinch J (1991) Perturbation methods. Cambridge Texts in Applied Mathematics. Cambridge University Press, Cambridge

    Google Scholar 

  20. Kevorkian J, Cole J (1996) Multiple scale and singular perturbation methods. Applied Mathematical Sciences, vol 114. Springer, New York

    Book  MATH  Google Scholar 

  21. Lagerstrom PA (1988) Matched asymptotic expansions. Applied Mathematical Sciences, vol 76. Springer-Verlag, New York

    Book  MATH  Google Scholar 

  22. Umemura A (1982) Matched-asymptotic analysis of low-Reynolds-number flow past two equal circular cylinders. J Fluid Mech 121:345–363

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Dorrepaal JM, O’Neil ME (1979) The existence of free eddies in a streaming stokes flow. Q J Mech Appl Math 32:95–107

    Article  MathSciNet  MATH  Google Scholar 

  24. Miyazaki T, Hasimoto H (1980) Separation of creeping flow past two circular cylinders. J Phys Soc Jpn 49(4):1611–1618

    Article  ADS  Google Scholar 

  25. Navier CLMH (1823) Mémoire sur les lois du mouvement des fluides. Mémoire de l’Académie Royale des Sciences de l’Institut de France, pp 389–440

  26. Skinner LA (1973) Generalized expansions for slow flow past a cylinder. Q J Mech Appl Mech 28(3):333–340

    Article  MATH  Google Scholar 

  27. Sen S, Mittal S, Biswas G (2009) Steady separated flow past a circular cylinder at low Reynolds numbers. J Fluid Mech 620:89–119

    Article  ADS  MATH  Google Scholar 

  28. Ranger KB (1995) Explicit solutions of the steady two-dimensional Navier–Stokes equations. Stud Appl Math 94(2):169–181

    Article  MathSciNet  MATH  Google Scholar 

  29. Greengard L, Kropinski MC, Mayo A (1996) Integral equation methods for Stokes flow and isotropic elasticity in the plane. J Comput Phys 125(2):403–414

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Imai I (1951) On the asymptotic behaviour of viscous fluid flow at a great distance from a cylindrical body, with special reference to Filon’s paradox. Proc R Soc Lond Ser A 208:487–516

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Kim S, Russel WB (1985) Modelling of porous media by renormalization of the Stokes equations. J Fluid Mech 154:269–286

    Article  ADS  MATH  Google Scholar 

  32. Vafai K, Kim SJ (1995) On the limitations of the Brinkman–Forchheimer-extended Darcy equation. Int J Heat Fluid Flow 16:11–15

    Article  Google Scholar 

  33. Jeffrey GB (1922) The rotation of two circular cylinders in a viscous fluid. Proc R Soc Lond Ser A 101:169–174

    Article  ADS  Google Scholar 

Download references

Acknowledgments

M.J.W. is grateful for the grant support by NSERC (Canada).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael J. Ward.

Appendices

Appendix 1: Drag coefficient

In [30], Imai gives an expression for the drag force in terms of an arbitrary closed circular contour of radius \(r_0\) surrounding the circular cross section of a cylinder. Converting this expression to polar coordinates, and defining the Reynolds number as \({\displaystyle \varepsilon }\equiv \text{ Re }\equiv {U_{\infty } L\rho _\mathrm{{f}}/\mu }\), we can express the drag coefficient, \(C_\mathrm{{D}}\), as

$$\begin{aligned} C_\mathrm{{D}}&= \frac{r_0}{2}\int \limits _{0}^{2\pi } \left[ \cos \theta \left( \frac{\partial \psi }{\partial r}\right) ^{2} - \frac{\cos \theta }{r^{2}}\left( \frac{\partial \psi }{\partial \theta } \right) ^{2}-\frac{2\sin \theta }{r}\frac{\partial \psi }{\partial r}\frac{\partial \psi }{\partial \theta }\right] {\Bigg \vert }_{r=r_0} \,\mathrm{{d}}\theta \nonumber \\&\quad -r_0\int \limits _{0}^{2\pi }(\sin \theta )\,\omega \frac{\partial \psi }{\partial \theta } {\Big \vert }_{r=r_0} \, \mathrm{{d}}\theta +\frac{r_0^{2}}{{\displaystyle \varepsilon }}\int \limits _{0}^{2 \pi }(\sin \theta )\frac{\partial \omega }{\partial r}{\Big \vert }_{r=r_0} \, \mathrm{{d}}\theta - \frac{r_0}{{\displaystyle \varepsilon }}\int \limits _{0}^{2\pi }(\sin \theta ) \, \omega {\vert }_{r=r_0} \,\mathrm{{d}}\theta . \end{aligned}$$
(7.12)

Here \(\psi \) satisfies (2.1)–(2.3), and the vorticity \(\omega \) is \(\omega = - \Delta \psi \).

To evaluate the integrals in this expression for \(C_\mathrm{{D}}\) when \({\displaystyle \varepsilon }\rightarrow 0\), we use the far-field behavior

$$\begin{aligned} \psi \sim S \left( r\log {r} - \frac{r}{2} + \frac{1}{2r}\right) \sin \theta , \end{aligned}$$
(7.13)

valid for \(r\rightarrow \infty \), of the Stokes solution. By symmetry, only the last two integrals in the expression for \(C_\mathrm{{D}}\) are nonzero, and consequently

$$\begin{aligned} C_\mathrm{{D}}\sim \frac{r_0^2}{{\displaystyle \varepsilon }}\int \limits _0^{2\pi } (\sin \theta )\, \frac{\partial \omega }{\partial r} {\Big \vert }_{r=r_0}\, \mathrm{{d}}\theta - \frac{r_0}{{\displaystyle \varepsilon }}\int \limits _0^{2\pi } (\sin \theta ) \omega \vert _{r=r_0}\, \mathrm{{d}}\theta . \end{aligned}$$
(7.14)

Since \(\omega =-{2Sr^{-1}}\sin \theta \), (7.14) becomes \(C_\mathrm{{D}}\sim 4\pi {\displaystyle \varepsilon }^{-1}S\), as given in (2.25).

Appendix 2: Stream function for flow past a porous cylinder

In this appendix we derive (5.1)–(5.6) for the stream function for flow past a porous cylinder. Inside the cylinder, we assume that the flow is governed by the Brinkman equation

$$\begin{aligned} \nabla P^i= -\frac{\mu }{k} \mathbf{v}^i + \hat{\mu } \Delta \mathbf{v}^i. \end{aligned}$$
(7.15)

Here \(k\) is the permeability of the porous cylinder, \(\mu \) is the dynamic viscosity of the fluid, \(\hat{\mu }\) is the effective viscosity of the porous medium, \(P^i\) is the pressure, and \(\mathbf{v^i=(v_r^i,v_{\theta }^i)}\) is the fluid velocity in the porous medium. The components of the fluid velocity in terms of stream function can be written as \(\mathbf{v}^i=(v_r^i,v_{\theta }^i)= (\partial \psi ^i/(r \partial \theta ), -\partial \psi ^i/ \partial r)\).

In general, the effective viscosity \(\hat{\mu }\) is not expected to be the same as the viscosity of the fluid \(\mu \). A derivation of the the dependency of the effective viscosity and permeability on the solid volume fraction in a porous medium through a renormalization of the Stokes equation was obtained in [31]. There, the authors showed that the Brinkman equation is valid in the limit of dilute porous media. In this limit, the effective viscosity, \(\hat{\mu }\), can be well approximated by the fluid viscosity, \(\mu \). Hence, in our derivation below, we assume that \(\hat{\mu } =\mu \).

We adopt the following scalings for Eq. (7.15):

$$\begin{aligned} \bar{P^i}= \frac{P^i}{\mu U_{\infty }/L},\quad \bar{\mathbf{v}^\mathbf{i}}= \frac{\mathbf{v}^\mathbf{i}}{U_{\infty }},\quad \bar{r}= \frac{r}{L}. \end{aligned}$$

Upon dropping the overbar, we write

$$\begin{aligned} \nabla P^i = -\chi ^2 \mathbf{v^i} + \Delta \mathbf{v^i}. \end{aligned}$$
(7.16)

Note that in the case where \(\hat{\mu } \ne \mu \), the preceding equation can be replaced by \(1/m\nabla P^i= -(\chi ^2/m)\mathbf{v^i} + \Delta \mathbf{v^i}\), where \(m=\hat{\mu }/\mu \). Therefore, the scaling, governing equations, and boundary conditions will not change. In (7.16), the pressure can be eliminated by taking the curl of the Brinkman equation. Then we obtain (5.2) upon substituting the velocity components in terms of the stream function. Outside the porous cylinder, the flow is governed by (5.1). Modeling the porous medium by the Brinkman equation rather than Darcy’s law allows us to satisfy the continuity of velocity and stresses at the interface between the fluid region and porous cylinder. The validity of the interfacial boundary conditions is discussed in [32]. We may write the boundary conditions at the interface as follows:

Continuity of velocity components

$$\begin{aligned} v_r^i&= v_r^e \quad \Rightarrow \quad \frac{\partial \psi ^i}{\partial \theta }= \frac{\partial \psi ^e}{\partial \theta } \Rightarrow \psi ^i=\psi ^e \quad \text{ on } \ r=1,\\ v_{\theta }^i&= v_{\theta }^e \quad \Rightarrow \quad \frac{\partial \psi ^i}{\partial r}= \frac{\partial \psi ^e}{\partial r} \quad \text{ on } \ r=1; \end{aligned}$$

Continuity of traction

$$\begin{aligned}&\textit{Normal stress balance:}\; [\![ n\cdot T\cdot n]\!]=0\;\text{ across } \quad r=1,\\&\textit{Tangential stress balance:}\; [\![n\cdot T\cdot t]\!]=0\;\text{ across } \quad r=1, \end{aligned}$$

where \([\![A]\!]=[A]^i-[A]^e\) denotes the jump in \(A\) across the surface of the porous cylinder. Here \(T\) is the traction tensor, and \(n=e_r\) and \(t=e_{\theta }\) are the unit normal and tangential vectors to the surface of the porous cylinder, respectively. We may write the traction tensor as follows:

$$\begin{aligned} T= \left( \begin{array}{cc} \tau _{rr}-P &{}\quad \tau _{r \theta } \\ \tau _{r \theta } &{}\quad \tau _{\theta \theta }-P \end{array}\right) =\left( \begin{array}{cc} 2\frac{\partial v_r}{\partial r} -P &{}\quad \frac{1}{r}\frac{\partial v_r }{\partial \theta }+r \frac{\partial }{\partial r} (\frac{v_{\theta }}{r})\\ \frac{1}{r}\frac{\partial v_r}{\partial \theta } +r \frac{\partial }{\partial r} (\frac{v_{\theta }}{r}) &{}\quad \frac{2}{r}\frac{\partial v_{\theta }}{\partial \theta } +\frac{2v_r}{r}-P \end{array}\right) . \end{aligned}$$

Upon writing the tangential and normal stress balance in terms of the stream function, we apply the continuity of velocity across the interface \(r=1\) to obtain

$$\begin{aligned}&\Big [\!\Big [\tau _{r\theta }\Big ]\!\Big ] = 0 \quad \Rightarrow \quad \Big [\!\Big [-\frac{1}{r^2}\frac{\partial ^2 \psi }{\partial \theta ^2}-\frac{1}{r}\frac{\partial \psi }{\partial r}+\frac{\partial ^2 \psi }{\partial r^2}\Big ]\!\Big ]=0 \quad \Rightarrow \quad \Big [\!\Big [{\partial ^2 \psi }/{\partial r^2}\Big ]\!\Big ]=0\\&\Big [\!\Big [\tau _{rr}-P\Big ]\!\Big ] = 0 \quad \Rightarrow \quad \Big [\! \Big [\frac{2}{r^2}\frac{\partial \psi }{\partial \theta }-\frac{2}{r} \frac{\partial ^2 \psi }{\partial r\partial \theta }-P\Big ]\!\Big ]=0 \quad \Rightarrow \quad \Big [\!\Big [P\Big ]\!\Big ]=0 \quad \Rightarrow \quad \Big [\!\Big [\frac{\partial P}{\partial \theta }\Big ]\!\Big ]=0. \end{aligned}$$

Then, upon using the Brinkman and Navier–Stokes equations inside and outside the porous cylinder, we obtain

$$\begin{aligned} -\frac{1}{r} \frac{\partial P^i}{\partial \theta }=-(\nabla ^2-\chi ^2) v_{\theta }^i=-\frac{1}{r}\frac{\partial }{\partial r}\left( r \frac{\partial v_{\theta }^i}{\partial r}\right) -\frac{1}{r^2} \frac{\partial ^2v_{\theta }^i}{\partial \theta ^2}-\frac{2}{r^2} \frac{\partial v_r^i}{\partial \theta }+\frac{v_{\theta }^i}{r^2}+ \chi ^2 v_{\theta }^i \end{aligned}$$

and

$$\begin{aligned} -\frac{1}{r} \frac{\partial P^e}{\partial \theta }=-\frac{1}{r} \frac{\partial }{\partial r}\left( r\frac{\partial v_{\theta }^e}{\partial r}\right) -\frac{1}{r^2}\frac{\partial ^2 v_{\theta }^e}{\partial \theta ^2}-\frac{2}{r^2}\frac{\partial v_r^e}{\partial \theta }+\frac{v_{\theta }^e}{r^2}\,. \end{aligned}$$

Upon equating these two expressions and then applying the continuity of velocity and tangential stress, we obtain that

$$\begin{aligned} -\frac{1}{r}\frac{\partial }{\partial r}\left( r\frac{\partial v_{\theta }^i}{\partial r}\right) +\chi ^2 v_{\theta }^i=-\frac{1}{r}\frac{\partial }{\partial r}\left( r\frac{\partial v_{\theta }^e}{\partial r}\right) \quad \Rightarrow \quad \frac{\partial ^3 \psi ^e}{\partial r^3}-\frac{\partial ^3 \psi ^i}{\partial r^3}=-\chi ^2\frac{\partial \psi ^i}{\partial r}\quad \text{ on }\; r=1. \end{aligned}$$

This completes the derivation of (5.1)–(5.6).

Appendix 3: Solution to Stokes problem for flow past two cylinders

In this appendix we outline the derivation of (6.12)–(6.16) using the formulation of [22]. Let \(\mathbf{y}=(y_1,y_2)^\mathrm{{t}}=r(\cos \theta ,\sin \theta )\), and introduce the bipolar coordinate system \((\xi ,\eta )\) defined by

$$\begin{aligned} y_1 = r \cos \theta = \frac{c\sinh \xi }{\cosh \xi -\cos \eta }, \quad y_2 = r \sin \theta = \frac{c\sin \eta }{\cosh \xi -\cos \eta }. \end{aligned}$$
(7.17)

Then lines of constant \(\xi \) map to the circles \((y_1-c\coth \xi )^2+y_2^2 = c^2 \text{ csch }^{2}\xi \). Therefore, if we choose \(\alpha \) and \(c\) by

$$\begin{aligned} c \coth \alpha ={d/2}, \quad c \, \text{ csch }\alpha = 1, \end{aligned}$$
(7.18)

so that

$$\begin{aligned} \alpha = \log \left( \frac{d}{2} + \sqrt{ \frac{d^2}{4}-1}\,\right) , \quad c= \sqrt{ \frac{d^2}{4} -1 }, \end{aligned}$$
(7.19)

it follows that the region \({\mathbb {R}}^2 \backslash D\) in (6.6) maps one to one to the rectangle \(\varOmega \equiv \lbrace {(\xi ,\eta ) \, \vert \,\,|\xi |\le \alpha , \, -\pi <\eta \le \pi \rbrace }\), and that \((\xi ,\eta )\rightarrow \mathbf{0}\) corresponds to \(r\rightarrow \infty \). For \(r\rightarrow \infty \), we readily obtain the following local behavior of the mapping (7.17):

$$\begin{aligned} \xi ^2 + \eta ^2 \sim \frac{4c^2}{r^2}, \quad \xi \sim \frac{2c \cos \theta }{r}, \quad \eta \sim \frac{2c \sin \theta }{r}. \end{aligned}$$
(7.20)

In terms of these bipolar coordinates, (6.6)–(6.9) is transformed into a problem for \(\varPhi (\xi ,\eta )\) given by (cf. [33])

$$\begin{aligned} \left[ \partial _{\xi \xi \xi \xi } + 2\partial _{\xi \xi \eta \eta } + \partial _{\eta \eta \eta \eta } - 2\partial _{\xi \xi } +2\partial _{\eta \eta } + 1\right] \varPhi = 0, \qquad (\xi ,\eta )\in \varOmega , \end{aligned}$$
(7.21)
$$\begin{aligned} \varPhi =\varPhi _{\xi } = 0, \quad \text{ on } \quad \xi =\pm \alpha ; \quad \varPhi \quad \text{ is } \,\, 2\pi \,\, \text{ periodic } \text{ in } \eta , \end{aligned}$$
(7.22)

with \(\varPhi \) having an appropriate singularity behavior as \((\xi ,\eta )\rightarrow \mathbf{0}\). In particular, \(\varPhi \rightarrow 0\) but is not \(C^1\) as \((\xi ,\eta )\rightarrow \mathbf{0}\). In terms of \(\varPhi \), the solution \(\psi _c\) to (6.6)–(6.9) is

$$\begin{aligned} \psi _c(y_1,y_2)=\psi \left[ \xi (y_1,y_2),\eta (y_1,y_2)\right] , \quad \text{ where } \; \psi (\xi ,\eta )= f \varPhi (\xi ,\eta ), \; f\equiv \frac{c}{\cosh \xi -\cos \eta }. \end{aligned}$$
(7.23)

The appropriate singularity behavior of \(\varPhi \) as \((\xi ,\eta )\rightarrow \mathbf{0}\) must be such that, in terms of polar coordinates, it yields \(\psi \sim r\log {r}\sin \theta \) as \(r\rightarrow \infty \). Since \(f\sim {2c/(\xi ^2+\eta ^2)}= {\mathcal O}(r^2)\) as \((\xi ,\eta )\rightarrow \mathbf{0}\) or, equivalently, as \(r\rightarrow \infty \), from (7.23) we conclude that \(\psi \) has the required far-field behavior provided that

$$\begin{aligned} \varPhi \rightarrow 0, \quad \text{ as } \quad (\xi ,\eta )\rightarrow \mathbf{0} \, \quad \text{ with } \text{ local } \text{ behavior } \quad \varPhi \sim -\frac{1}{2}\eta \log (\xi ^2 + \eta ^2). \end{aligned}$$
(7.24)

To verify this claim, we calculate for \((\xi ,\eta )\rightarrow \mathbf{0}\) using (7.20) that

$$\begin{aligned} \psi =f \varPhi \sim \frac{2c}{\xi ^2+\eta ^2}\left( -\frac{\eta }{2} \log (\xi ^2+\eta ^2)\right) \sim \frac{2c r^2}{4c^2}\left( -\frac{1}{2} \frac{2c\sin \theta }{r} \log \left( \frac{4c^2}{r^2}\right) \right) \sim r\log {r}\sin \theta \end{aligned}$$
(7.25)

as \(r\rightarrow \infty \). Therefore, problem (6.6)–(6.9) is transformed into a problem of finding a solution to (7.21)–(7.22) satisfying the local behavior in (7.24) as \((\xi ,\eta )\rightarrow \mathbf{0}\).

Next, we construct an exact solution of (7.21) with the local behavior (7.24). To determine such a solution, we recall in terms of polar coordinates that \(\Delta ^2 \left[ \frac{r^2}{2}\log {r}\right] =0\). Upon differentiating with respect to \(y_2\), we obtain that \(\Delta ^2[ y_2\log {r} + {y_2/2}]=0\), so that \(\Delta ^2[y_2\log {r}]=0\). To convert this exact biharmonic solution having the correct behavior at infinity to bipolar coordinates, we first calculate \(\log {r}\) from (7.17) as

$$\begin{aligned} \log {r} = \log \left( \frac{c\left( \sinh ^2\xi +\sin ^2\eta \right) ^{1/2}}{{\cosh \xi -\cos \eta }}\right) = -\frac{1}{2}\log \left( \cosh \xi - \cos \eta \right) + \frac{1}{2}\log \left( \cosh \xi + \cos \eta \right) + \log {c}. \end{aligned}$$
(7.26)

Then, upon calculating \(y_2\log {r}\) from using (7.17) for \(y_2\), it follows that an exact solution to (7.21) with the local behavior (7.24) is

$$\begin{aligned} \tilde{\varPhi }_0 = -\frac{1}{2}\sin \eta \log \left( \cosh \xi - \cos \eta \right) + \frac{1}{2}\sin \eta \log \left( \cosh \xi + \cos \eta \right) . \end{aligned}$$
(7.27)

The second term in (7.27) is regular as \((\xi ,\eta )\rightarrow \mathbf{0}\) but is not \(C^{1}\) at \((\xi ,\eta )=(0,\pm \pi )\). Moreover, by expanding this term in a Fourier sine series in \(\eta \) to generate a separation-of-variables type of solution, it is readily verified that it satisfies (7.21). As such, an exact solution to (7.21), which has the local behavior (7.24) as \((\xi ,\eta )\rightarrow \mathbf{0}\) but is otherwise smooth in \(\varOmega \), is simply the first term in (7.27), which we label by

$$\begin{aligned} \varPhi _0\equiv -\frac{1}{2}\sin \eta \log \left( \cosh \xi -\cos \eta \right) . \end{aligned}$$
(7.28)

We then must add to \(\varPhi _0\) a smooth solution \(\varPhi _s\), constructed using the separation of variables applied to (7.21), in order to satisfy the boundary conditions in (7.22) at \(\xi =\pm \alpha \). Since \(\varPhi _0\) is even in \(\xi \), we need only choose \(\varPhi _s\) to be an even function of \(\xi \), with \(\varPhi _s=-\varPhi _0\) and \(\partial _\xi \varPhi _s=-\partial _\xi \varPhi _0\) on \(\xi =\alpha \).

To expand \(\varPhi _0\) as a Fourier sine series in \(\eta \), we first define \(z=\mathrm{{e}}^{-\xi +\mathrm{{i}}\eta }\) and derive the identity

$$\begin{aligned} \sum _{m=1}^{\infty }\frac{\mathrm{{e}}^{-m\xi }}{m}\cos (m\eta )=\mathrm{Re}\left[ \sum _{m=1}^{\infty } \frac{z^m}{m}\right] =-\log |1-z|= \frac{\xi }{2}-\frac{\log {2}}{2} - \frac{1}{2} \log \left( \cosh \xi -\cos \eta \right) . \end{aligned}$$
(7.29)

From this identity, and using trigonometric addition formulas, we obtain for \(\xi >0\) that \(\varPhi _0\) has the Fourier sine series

$$\begin{aligned} \varPhi _0&= -\frac{\sin \eta }{2}\left[ \xi -\log {2} -2\sum _{m=1}^{\infty } \frac{\mathrm{{e}}^{-m\xi }}{m} \cos (m\eta )\right] ,\nonumber \\&=-\frac{1}{2}\left( \xi -\log {2} + \frac{1}{2}\mathrm{{e}}^{-2\xi }\right) \sin (\eta ) + \frac{1}{2} \sum _{k=2}^{\infty } \left( \frac{\mathrm{{e}}^{-(k-1)\xi }}{k-1} - \frac{\mathrm{{e}}^{-(k+1)\xi }}{k+1} \right) \sin (k\eta ). \end{aligned}$$
(7.30)

Then, since the Fourier sine series separation of variables solution \(\varPhi _s\) for (7.21), which is even in \(\xi \), is a linear combination of \(\cosh [(k+1)\xi ]\sin (k\eta )\) and \(\cosh [(k-1)\xi )]\sin (k\eta )\) for \(k\ge 1\), it follows that the solution \(\varPhi =\varPhi _0+\varPhi _s\) to (7.21)–(7.23) with (7.24) has the form

$$\begin{aligned} \varPhi = h_1(\xi ) \sin (\eta ) + \sum _{k=1}^{\infty } h_{k}(\xi )\sin (k\eta ), \end{aligned}$$
(7.31)

where

$$\begin{aligned} h_1(\xi ) \equiv -\frac{\xi }{2} + \frac{\ln {2}}{2} - \frac{\mathrm{{e}}^{-2\xi }}{4} - \frac{a_1}{2}\cosh (2\xi ) - \frac{b_1}{2}, \end{aligned}$$
(7.32)
$$\begin{aligned} h_k(\xi ) \equiv \frac{1}{2}\left( \frac{\mathrm{{e}}^{-(k-1)\xi }}{k-1} - \frac{\mathrm{{e}}^{-(k+1)\xi }}{k+1} \right) - \frac{a_k}{2}\cosh [(k+1)\xi ] - \frac{b_k}{2} \cosh [(k-1)\xi ], \quad k\ge 2. \end{aligned}$$
(7.33)

Upon setting \(h_k=h_k^{\prime }=0\) at \(\xi =\alpha \) for \(k\ge 1\), we obtain that \(a_k\) and \(b_k\) for \(k\ge 1\) satisfy the linear system

$$\begin{aligned} a_1\cosh (2\xi ) + b_1=- \alpha + \ln {2} - \frac{1}{2}\mathrm{{e}}^{-2\alpha }, \quad 2a_1 \sinh (2\alpha )=-1 + \mathrm{{e}}^{-2\alpha },\end{aligned}$$
(7.34)
$$\begin{aligned} a_k \cosh \left[ (k+1)\alpha \right] + b_k \cosh \left[ (k-1)\alpha \right] = \frac{\mathrm{{e}}^{-(k-1)\alpha }}{k-1}- \frac{\mathrm{{e}}^{-(k+1)\alpha }}{k+1}, \quad k\ge 2,\end{aligned}$$
(7.35)
$$\begin{aligned} a_k (k+1) \sinh \left[ (k+1)\alpha \right] + b_k (k-1)\sinh \left[ (k-1) \alpha \right] = \mathrm{{e}}^{-(k+1)\alpha } - \mathrm{{e}}^{-(k-1)\alpha }, \quad k\ge 2. \end{aligned}$$
(7.36)

The solution of this system is given in (6.12)–(6.16).

Finally, we calculate the correction to the local behavior of \(\varPhi \) as \((\xi ,\eta )\rightarrow \mathbf{0}\). We obtain from (7.31)–(7.33) that

$$\begin{aligned} \varPhi \sim -\frac{\eta }{2}\log \left( \frac{\xi ^2+\eta ^2}{2}\right) - \frac{\eta }{2} \sum _{m=1}^{\infty } m (a_m + b_m) \quad \text{ as }\ (\xi ,\eta )\rightarrow \mathbf{0}. \end{aligned}$$
(7.37)

Recalling that \(\psi =f\varPhi \), we use the local behavior (7.20) to obtain that \(\psi _c\) has the form (6.9), where \(q\) is defined in (6.12). This completes the derivation of (6.12)–(6.16).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hormozi, S., Ward, M.J. A hybrid asymptotic-numerical method for calculating drag coefficients in 2-D low Reynolds number flows. J Eng Math 102, 3–33 (2017). https://doi.org/10.1007/s10665-014-9701-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10665-014-9701-x

Keywords

Navigation