16.4.7 Flow Rate

As for the other cross-sections we will now derive the flow rate for the rectangular cross-section according to Eq. 10.9. Obviously, any solution to the Navier-Stokes equation we found in Eq. 16.53, Eq. 16.58, and Eq. 16.62 could be used for deriving the flow rate. In the following, we will limit ourselves to Eq. 16.53 since this is the most commonly found equation, which gives rise to some convenient simplifications. Using Eq. 10.9 and Eq. 16.53 we find the flow rate as

Q=AudA=4h2ηπ3dpdxn=01(2n+1)3w2w2(1cosh((2n+1)πyh)cosh((2n+1)πw2h))dy0hsin((2n+1)πzh)dz=4h2ηπ3dpdxn=01(2n+1)3[(yh(2n+1)πsinh((2n+1)πyh)cosh((2n+1)πw2h))]w2w2[h(2n+1)πcos((2n+1)πzh)]0h=4h2ηπ3dpdxn=01(2n+1)3((w2+w2)(h(2n+1)πsinh((2n+1)πw2h)cosh((2n+1)πw2h)h(2n+1)πsinh((2n+1)πw2h)cosh((2n+1)πw2h)))(h(2n+1)π(1cos((2n+1)π)1))=4h2ηπ3dpdxn=01(2n+1)3(w2h(2n+1)πtanh((2n+1)πw2h))2h(2n+1)π=4h2ηπ3dpdxn=01(2n+1)42hπ(w2h(2n+1)πtanh((2n+1)πw2h))=8h3ηπ4dpdxn=0(w(2n+1)42h(2n+1)5πtanh((2n+1)πw2h))

si99_e  (Eq. 16.63)

We can further simplify this expression using the fact that Eq. 16.63 contains a series with a known convergence value given by

n=11(2n+1)4=196π4

si100_e

See Tab. 4.1 for details on this series. We can therefore rewrite Eq. 16.63 to

Q=h2w12ηdpdx(1192hπ5wn=01(2n+1)5tanh((2n+1)πw2h))

si101_e  (Eq. 16.64)

Eq. 16.64 gives the general case of the flow rate in a rectangular channel.

Channels With Square Cross-Sections. If setting h = w in Eq. 16.64, we obtain

Q=w212ηdpdx(1192π5n=01(2n+1)5tanh((2n+1)πw2))

si102_e  (Eq. 16.65)

Approximation for Channels With Low Aspect Ratios. As you can see from Fig. 3.4b the hyperbolic tangent tanh x converges to 1 for values of x > 1. Therefore, for channels with aspect ratios r si183_e 1 we will find

whtanh((2n+1)πw2h)1

si103_e

in which case we can simplify Eq. 16.64 to

Q=h3w12ηdpdx(1192hπ5wn=01(2n+1)5)

si104_e  (Eq. 16.66)

where we can apply a second known series convergence value given by

n=11(2n+1)5=3132ζ(5)1.0045

si105_e

Again, see Tab. 4.1 for details on this series. Using this series we can rewrite Eq. 16.66 to

Q=h3w12ηdpdx(1192hπ5w11.0045)=h3w12ηdpdx(10.63hw)

si106_e  (Eq. 16.67)

For hw0si107_e we obtain the flow rate calculated for the infinitesimally elongated channel, Eq. 16.14. Obviously, we make mistakes if applying Eq. 16.67 for channels with aspect ratios in the range of 1 and above. Fig. 16.15 shows the relative error we obtain when using Eq. 16.67 instead of Eq. 16.64 for calculating the flow rate. As you can see, the error is in the range of 13 % for r = 1, whereas it drops below 1 % at r < 0.65. For smaller aspect ratios, the error is negligible. For aspect ratios bigger than 1 the simplification should not be used since the error introduced is significant.

f16-15-9781455731411
Fig. 16.15 Simplification errors for the flow rate when using Eq. 16.67 instead of Eq. 16.64 as a function of the aspect ratio r.

16.5 Summary

In this section we have studied analytical solutions to Poiseuille flow problems, i.e., flows that are pressure-driven. We have derived solutions to several commonly encountered flow channel geometries in microfluidics, including the elliptical and circular as well as the rectangular channel profile. For these profiles we have derived the velocity profile as well as the flow rate equation. As we have seen, even a simple cross-section such as a rectangular channel required the application of concepts for solving PDEs that we have derived in section 8.2 and section 8.3. In most cases, we were able to work with very simple Dirichlet-type boundary conditions because the flow velocity on the solid wall of a microfluidic channel is 0. Until now, we have not yet solved more complex flow conditions where one of these boundary conditions may not be zero as in the case of, e.g., the Couette flow. If we want to solve flow profiles in channels with changing cross-section or more complex geometries, we will very quickly run out of analytical techniques for assessing such complex flow cases. This is the regime of numerical solutions to the Navier-Stokes equation, which we will explore in section 28.

References

[1] Hagen G. “Über die Bewegung des Wassers in engen zylindrischen Röhren.”. Pogg. Ann. 1839;46:423–442 (cit. on p. 329).

[2] Darcy H. Les fontaines publiques de la ville de Dijon. Exposition et application à suivre et des formules à employer dans les questions de duistribution d’eau. 1856 (cit. on p. 334).

[3] Darcy H. . Recherches expérimentales relatives au mouvement de l’eau dans les tuyaux. 1857;Vol. 1. Mallet-Bachelier, (cit. on p. 334).

[4] Simmons C.T. Henry Darcy (1803–1858): Immortalised by his scientific legacy. Hydrogeology Journal. 2008;16(6):1023–1038 (cit. on p. 334).

[5] Colebrook C.F. Turbulent Flow in Pipes, with particular reference to the Transition Region between the Smooth and Rough Pipe Laws. Journal of the Institution of Civil Engineers. 1939;11(4):133–156 (cit. on p. 335).

[6] Colebrook C.F., White C.M. Experiments with Fluid Friction in Roughened Pipes. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. 1937;161(906):367–381 (cit. on p. 335).

[7] Rollmann P., Spindler K. Explicit representation of the implicit Colebrook–White equation. Case Studies in Thermal Engineering. 2015;5:41–47 (cit. on p. 335).


1 Gotthilf Hagen was a German engineer who studied the flow of fluids in small capillaries independently of Jean Poiseuille. Hagen published his results one year before Poiseuille in 1839 [1]. Hagen and Poiseuille derived the fluid mechanics correctly and formulated the equation for calculating the total flow through the profile as well as the pressure drop, although Hagen derived the flow profile incorrectly.

1 Henry Darcy was a French engineer who made important contributions to the fluid mechanics of hydraulic systems. In his most important work Darcy described (under the well-hidden Note D) his famous experiments on sand columns from which he derived Darcy’s law. This work is sometimes referred to as the birth of hydrology [2]. Darcy also provided the correlation that would be known as the Darcy friction coefficient, f[3]. This equation was a refined version of an earlier experimental correlation provided by French mathematician Gaspard de Prony. A more detailed account on the work and life of Darcy can be found in the noteworthy paper by Simmons [4].

2 Julius Weisbach was a German mathematician who made important contributions to the application of mathematics for engineering mechanical problems. He may be most renowned for his contribution to the Darcy-Weisbach equation, as he was the first to propose the equation in its most common form.

1 Cyril Colebrook was a British physicist who made important contributions to fluid mechanics. Among his most important work are studies on turbulent flows in pipes. He most notably derived the Colebrook-White equation, which is used to approximate the Darcy friction coefficient [5].

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset