Chapter 16

Analytical Solutions to Poiseuille Flow Problems in Different Geometries

16.1 Elliptical and Circular Profiles

16.1.1 Analytical Solution: An Approach

Circular profiles are a special case of elliptical profiles, which is why we will treat them together. Obviously, these profiles are very relevant to microfluidics as they represent the case of a circular tube or a squeezed capillary in the case of the elliptical profile.

16.1.1.1 Approach for vx

We do not have an intuitive approach for vx, which is why we first default to the ellipses equation that is given by

(xa)2+(yb)2=1

si2_e

This function simply describes the contour of our cross-section, where a and b are the ellipse’s half-axis along the y-axis and the half-axis along the z-axis, respectively (see Fig. 16.1a). Obviously, for a = b = r we obtain the equation for a circle.

f16-01-9781455731411
Fig. 16.1 Schematic of the Poiseuille flow in an elliptical cross-section in Cartesian coordinates (a) and polar coordinates (b).

As stated, we assume no-slip condition for Poiseuille flows, which is why a good first guess would be to choose a function that would ensure that the velocity is in fact zero on the contour ∂C. In this case our approach would read

vx(y,z)=v0(1(ya)2+(zb)2)

si3_e  (Eq. 16.1)

This function would at least ensure that for all values (x, y) on the contour, we would obtain a velocity of 0. However, we introduced an unknown variable here, namely v0. If our approach of vx is correct we should be able to determine this unknown variable after solving the PDE.

16.1.1.2 Verification

Now we insert Eq. 16.1 into Eq. 15.14 and set kx = 0. In this case we find

dpdx=η(2vxz2+2vxy2)=ηv0(2z2(1(ya)2+(zb)2)+2y2(1(ya)2+(zb)2))=ηv0(2z2((zb)2+2y2(ya)2))=2ηv0(1b2+1a2)=2ηv0a2+b2a2b2v0=12ηdpdxa2b2a2+b2

si4_e  (Eq. 16.2)

This looks surprisingly promising. Our approach actually solved the PDE and we were able to determine the unknown variable v0.

16.1.2 Velocity Profile for Elliptical Cross-Sections

We are now able to describe the velocity profile using our approach and Eq. 16.2. We find

vx(y,z)=12ηdpdxa2b2a2+b2(1(ya)2+(zb)2)

si5_e  (Eq. 16.3)

Eq. 16.3 is the equation for the velocity profile of the ellipse. Please note that this equation suggests that the velocity profile is actually negative. This is not the case: The pressure drop dpdxsi6_e is also negative as the flow is driven from areas of higher pressure to areas of lower pressure. In many textbooks you may see the pressure gradient being replaced by a pressure drop over a given length of the channel Δplsi7_e, in which case Eq. 16.3 can be rewritten to

vx(y,z)=Δp2ηla2b2a2+b2(1(ya)2+(zb)2)

si8_e  (Eq. 16.4)

16.1.3 Flow Rate for Elliptical Cross-Sections

After having derived the velocity profiles we will now derive the flow rate for the Poiseuille flow in cylindrical coordinates according to Eq. 10.9 as

dVdt=AvdA

si9_e

As the cross-section is elliptical we will carry out this integral in polar coordinates (see Fig. 16.1b). For this we need to transform Eq. 16.4 into polar coordinates. We will express y and z as function of the cylindrical radius 0 ≤ ρ ≤ 1 and the angle 0 ≤ φ ≤ 2π, resulting in

y(ρ,φ)=ρacosφz(ρ,φ)=ρbsinφ(ya)2+(zb)2=ρ2cos2φ+ρ2sin2φ=ρ2vx(ρ,φ)=12ηdpdxa2b2a2+b2(1ρ2)

si10_e

In order to transform the integral from Cartesian to polar coordinates, we need to find the Jacobian determinant (see section 7.5) given by

detJ=|yρyφzρzρ|=|ρ(ρacosφ)ρ(ρacosφ)ρ(ρbsinφ)ρ(ρbsinφ)|=|acosφρasinφbsinφρbcosφ|=acosφρbcosφ+ρasinφbsinφ=ρab(sin2φ+cos2φ)=ρab

si11_e

Having derived the determinant we can now carry out the integration:

dVdt=0102πv(ρ,φ)ρabdφdρ=12ηdpdxa3b3a2+b201(1ρ2)ρdρ02πdφ=12ηdpdxa3b3a2+b2[ρ22ρ44]012πQ=π4ηdpdxa3b3a2+b2

si12_e  (Eq. 16.5)

For details on how the Jacobian determinant is used for integral transformations refer to section 7.3.

16.1.4 Velocity Profile for Circular Cross-Sections

As stated, by simply replacing a = b = r we obtain the velocity profile for the circle from Eq. 16.3:

vx(y,z)=12ηdpdxr22(1(yr)2+(zr)2)=14ηdpdx(r2(y2+z2))

si13_e  (Eq. 16.6)

or

vx(y,z)=Δ4ηldpdx(r2(y2+z2))

si14_e  (Eq. 16.7)

if you prefer the notation with the pressure drop over the length of the channel.

16.1.5 Flow Rate for Circular Cross-Sections

Applying a = b = r to Eq. 16.5 yields the flow rate for cylindrical cross-sections given by

dVdt=π4ηdpdxr22r2Q=πr48ηdpdx

si15_e  (Eq. 16.8)

16.1.6 Visualization

In the following we will visualize the equations we found using the Maple listing shown in listing 16.1.

Listing 16.1

[Maple] Listing for the profile of elliptical Poiseuille flows. A digital version of the listing can be found under the name PoiseuilleFlowEllipticalCircular.mw in the supplementary material.

1   restart:read "Core.txt":
2
3   #profile for the ellipsis
4   velx_ellipse:=(y,z,dpdx,a,b,eta)-> -1/(2*eta)*dpdx*(a^2*b^2)/(a^2+b^2)*(1 - ((y/a)^2 + (z/b)^2));
5
6   #we define a helper function that takes care of the units
7   velx_ellipse2:=(y,z,dpdx,a,b)-> simplify(velx_ellipse(y*Unit(um),z*Unit(um),dpdx*Unit(mbar/mm),a*Unit(um),b*Unit(um),1*Unit(mPa*s))/Unit(mm/s));
8
9   #now plot the profile for an elliptical channel
10     quickplot3d([velx_ellipse2(y,z,-0.001,100,50),velx_ellipse2(y,z,-0.01,100,50),velx_ellipse2(y,z,-0.02,100,50)],y=-100..100,z=0..50,0..2,"y [um]","z [um]",typeset(v[x]^empty," [mm/s]"));
11
12  #now plot the profile for a round channel
13     quickplot3d([velx_ellipse2(y,z,-0.001,100,100),velx_ellipse2(y,z,-0.01,100,100),velx_ellipse2(y,z,-0.02,100,100)],y=-100..100,z=0..100,0..2,"y [um]","z [um]",typeset(v[x]^empty," [mm/s]"));

In line 1 we restart the Maple server and include Core.txt. In line 4 we define the flow profile. This is Eq. 16.3, which is called velx_ellipse in the listing. In order to simplify working with units, we define a second function velx_ellipse2, which passes on the independent variables with suitable units. We choose µ m for the dimensions and mbar mm−1 for the pressure drop. Also, we pass along the viscosity value of water. If you want to describe the flow for different fluids, change the variable eta.

Finally in line 10 we calculate the velocity profiles for an elliptical channel with a = 100 µ m and b = 50 µ m. We use the function quickplot3d for creating the plot. Refer to section 2.3.3 for details about this function. We supply different values for the pressure drop, i.e., − 0.001 mbar mm−1, − 0.01 mbar mm−1, and − 0.02 mbar mm−1. The output of the function is shown in Fig. 16.2a. As you can see, for the last value, the flow velocity in the center of the channel approaches 2 mm s−1.

f16-02-9781455731411
Fig. 16.2 Calculated Poiseuille flow profiles for elliptical and circular channel cross-sections. The assumed pressure drops were − 0.001 mbar mm−1 (dark blue), − 0.01 mbar mm−1 (medium green), and − 0.02 mbar mm−1 (light yellow). The fluid used was water.

The second case we study is the flow in a circular tube. Here, we use the same values for the different pressure drops and assume a channel of a = b = r = 50 µ m. For the highest pressure drop, the center flow velocity is around 1.25 mm s−1. The calculated flow velocity profiles are shown in Fig. 16.2b.

16.2 Planar Infinitesimally Extended Channel Cross-Sections

A flow profile that resembles the Hagen-Poiseuille flow profile (see section 16.3) is the Poiseuille flow in planar, infinitesimally extended channels. This is a special case of the a Poiseuille flow because it is effectively a one-dimensional problem. In this it resembles the Hagen-Poiseuille profile. As we will see, the Hagen-Poiseuille flow profile is a one-dimensional fluid mechanical problem since it exploits the symmetry of a circular tube. The flow in infinitesimally extended channels is one-dimensional because we assume all changes in one direction (in this case along the y-axis) to be 0 (see Fig. 16.3). Fig. 16.3b gives an indication how this profile stretches in the channel and why we can consider it a one-dimensional problem.

f16-03-9781455731411
Fig. 16.3 Poiseuille flow in a planar infinitesimally extended channel. a) Flow profile, b) flow profile as formed in the channel, c) shear stress profile.

16.2.1 Derivation

We again start with Eq. 15.14, reducing the equation to the one-dimensional case for which we find

dpdx=η2vxz2dpdx=ηd2vxdz2

si16_e  (Eq. 16.9)

Here we have also neglected the volume forces setting kx = 0. Eq. 16.9 is a second-order ODE for which we need two boundary conditions, which are given by the no-slip condition at the two channel walls: vx (h/2) = 0, vx (−h/2) = 0. We integrate Eq. 16.9 twice finding

dvxdz=1ηdpdxz+c1vx(z)=12ηdpdxz2+c1z+c2

si17_e  (Eq. 16.10)

Inserting the boundary conditions results in

0=012ηdpdxh24+c1h2+c20=012ηdpdxh24c1h2+c2

si18_e

from which can find

c2=12ηdpdxh24

si19_e  (Eq. 16.11)

c1=0

si20_e  (Eq. 16.12)

Inserting Eq. 16.11 and Eq. 16.12 into Eq. 16.10 yields

vx(z)=12ηdpdx(z2h24)vx(z)=12ηdpdx(z2(h4)2)vx(z)=h28ηdpdx(1(zh2)2)

si21_e  (Eq. 16.13)

16.2.2 Velocity Profile

Eq. 16.13 described the velocity profile for the planar infinitesimally extended channel. It is a parabola with the maximum velocity

vx,max=h28ηdpdx

si22_e

Again, the pressure difference is negative since we expect a pressure drop along the channel: dpdx<0si23_e. In this case, the overall velocity becomes positive.

16.2.3 Flow Rate

Having derived the velocity profile we can calculate the flow rate using Eq. 10.9 as

Q=AvdA=h28ηdpdx0h(1(zh2)2)ωdz=ωh28ηdpdx[(z43H2z3)]h2h2=ωh28ηdpdx(h2h6(h2h6))=ωh312ηdpdx

si24_e  (Eq. 16.14)

16.2.4 Shear Stress Profile

The shear stress profile can be obtained from Eq. 9.5 and Eq. 16.13. It is given by

τ(z)=h28dpdx2zh2=dpdxz

si25_e  (Eq. 16.15)

=dpdx|z|

si26_e  (Eq. 16.16)

The profile needs to be symmetric although Eq. 16.15 does not suggest it: For negative values of z you will obtain negative shear forces, which is physically not meaningful. This can be accounted for, e.g., by only considering the absolute value (see Eq. 16.16).

16.2.5 Visualization

Fig. 16.4 shows the velocity and the shear force profile in an infinitesimally extended channel with 10 µ m height. As you can see, the velocity profile is a parabola, whereas the shear force profile is linear. The maximum velocity in the channel center is about 0.125 mm s−1. Fig. 16.5 shows the calculated profiles for a channel with 50 µ m height. Here, the maximum flow velocity can reach up to above 1.5 mm s−1.

f16-04-9781455731411
Fig. 16.4 Velocity and shear force profile in a channel of 10 µm height.
f16-05-9781455731411
Fig. 16.5 Velocity and shear force profile in a channel of 50 µ m height.

16.3 Flows in Circular Cross-Sections: Hagen-Poiseuille Flow

16.3.1 Introduction

Obviously, circular cross-sections are a very commonly occurring in technical flow systems. Macroscopic examples include pipes and the tap; microfluidic examples include capillaries and circular tubes. The fluid mechanics of circular tubes or channels is commonly referred to as Hagen-Poiseuille1flow. As we will see, the flow profile is parabolic in shape and the shear stress is linear (see Fig. 16.6). Calculating the flow in circular cross-sections is not difficult as it exploits the fact that the Navier-Stokes equation in cylindrical coordinates is one-dimensional. There are two approaches deriving the same equation. We will first look at the direct solution of the Navier-Stokes equation in cylindrical coordinates which is the Euler equation.

f16-06-9781455731411
Fig. 16.6 Hagen-Poiseuille flow in a circular tube. Flow profile (a) and shear stress profile (b).

16.3.2 Derivation Using the Euler Equation

Derivation. The first method for approaching the fluid mechanics of circular tube can be directly derived from the Euler equation (see Eq. 14.8), which we derived for the circular flow tube. As such, a circular tube or pipe is nothing else than a circular flow tube, which makes this approach very intuitive. Again we only consider the fully-developed and stationary case and neglect gravity. In this case the Euler equation simplifies to

τr+τr=dpdz

si27_e  (Eq. 16.17)

where we assume a pressure drop along the x-axis, which is independent of r; this allows us to treat dpdxsi6_e as a constant. If we reduce the fluids to incompressible Newtonian fluids we can use Eq. 9.5 to obtain a dependency between the shear forces τ and the velocity v. Please note that, as we will be working with cylindrical coordinates, we use z as the coordinate axis along the channel, in which case Eq. 16.17 can be written as

ηrdvdr+ηd2vdr2=dpdz1rdvdr+d2vdr2=1ηdpdz

si30_e  (Eq. 16.18)

Analytical Solution. Eq. 16.18 is a second-order inhomogeneous ODE for which we need to find a suitable solution. Listing 16.2 shows the Maple code we can use to obtain the required solution.

Listing 16.2

[Maple] Listing for the mechanics of the tube flow. This flow case is referred to as the Hagen-Poiseuille flow. A digital version of the listing can be found under the name TubeFlow.mw in the supplementary material.

1    restart:read "Core.txt":
2    #we define the differential equation and the boundary values
3    ode1:=1/r*diff(v(r),r)+diff(diff(v(r),r),r) = 1/eta*dpdz;
4    bc1:=v(R)=0;bc2:=D(v)(0)=0;
5
6    #and solve the ODE
7    v:=unapply(rhs(dsolve({ode1,bc1,bc2},v(r))),r,R,eta,dpdz);
8
9    #we define a helper function that will take care of the units
10   v2:=(r,R,dpdz)->simplify(v(r*Unit(um),R*Unit(um),1*Unit(mPa*s),dpdz*Unit(mbar/mm))/Unit(mm/s)):
11
12   #we plot for a 10 um diameter tube and varying pressure drops
13   quickplot([r->v2(r,5,-0.01),r->v2(r,5,-0.05),r->v2(r,5,-0.1)],-5..5,"radius [um]","velocity [mm/s]");
14   #we plot for a 50 um diameter tube and varying pressure drops
15   quickplot([r->v2(r,25,-0.01),r->v2(r,25,-0.05),r->v2(r,25,-0.1)],-25..25,"radius [um]","velocity [mm/s]");
16
17   #calculate the shear forces
18   tau:=unapply(eta*diff(v(r,R,eta,dpdz),r),r,dpdx);
19
20   #again we define a little helper function
21   tau2:=(r,dpdz)->simplify(abs(tau(r*Unit(um),dpdz*Unit(mbar/mm))/Unit(N/m^2))):
22
23   #we plot for a 10 um diameter tube and varying pressure drops
24   quickplot([r->tau2(r,-0.01),r->tau2(r,-0.05),r->tau2(r,-0.1)],-5..5,"radius [um]",typeset("shear stress ",N/m^2));
25
26   #we plot for a 50 um diameter tube and varying pressure drops
27   quickplot([r->tau2(r,-0.01),r->tau2(r,-0.05),r->tau2(r,-0.1)],-25..25,"radius [um]",typeset("shear stress ",N/m^2));

The code is not very complicated. In line 1 we restart the Maple server and include Core.txt. In line 3 we define the ODE and the two boundary conditions. The first boundary condition is given by the fact that the velocity at the radius r = R must be zero (no-slip boundary condition). Furthermore, we expect a rotationally symmetric flow profile in which case the first derivative of the velocity profile at the center (r = 0) must also be zero. Please note that we used the variable dpdz to describe the pressure drop, which will make it easier to work with the resulting function that takes dpdz as an independent variable.

Finally, in line 7 we instruct Maple to solve the ODE. The function unapply allows us to directly assign the right-hand side of our resulting equation to the function v, which uses the independent variables r, the tube radius R, the fluid’s viscosity η, and the pressure drop dpdxsi6_e. Maple returns the following function:

v(r)=14ηdpdz(r2R2)=R24ηdpdz((rR)21)

si32_e  (Eq. 16.19)

=vmax((rR)21)

si33_e  (Eq. 16.20)

Velocity Profile. Eq. 16.20 is the Hagen-Poiseuille flow profile for the rotationally symmetric tube or pipe. It is a parabolic profile with the maximum velocity at the center of the tube. In order to visualize this equation, we define a helper function v2 in line 10. This new function will take the r and R in mm and dpdzsi34_e in mbar/mm−1. The viscosity value is set for the value of water. Finally, v2 is divided by mm s−1 and a simplification is applied to it. This results in the function outputting numerical values, which is a prerequisite to plotting functions in Maple.

Starting from line 13 the flow profile of Eq. 16.20 is calculated for two tube diameters using different values for dpdzsi34_e: − 0.01 mbar/mm−1, − 0.05 mbar/mm−1, and − 0.1 mbar/mm−1. Obviously all of these values must be negative since we assume the pressure to decrease as we move along z due to the losses caused by the fluid’s internal friction. Fig. 16.7 shows the two sets of profiles calculated for circular tubes with diameters 10 µ m and 50 µ m, respectively. As you can see, the flow velocities can become significant for a higher tube diameter R and increasing values of dpdzsi34_e.

f16-07-9781455731411
Fig. 16.7 Flow profiles calculated for two tubes with different diameters. The fluid used was water, and the pressure drop along the z-axis is given.

Flow Rate. Having derived the velocity profile we can now derive the flow rate according to

Q=AvdA

si37_e  (Eq. 16.21)

=R24ηdpdz0R((rR)21)2πrdr=πR22ηdpdzπ0R(r2R3r)dr=πR22ηdpdzπ[r44R2r22]0R=πR48ηdpdz

si38_e  (Eq. 16.22)

where we have used the infinitesimal area of the circle dA = 2 π r dr (see Eq. 3.104). Eq. 16.22 is sometimes referred to as the Hagen-Poiseuille law. It is a very important function for calculating the flow in technical system, such as pipe and hydraulic systems. From all the equations commonly encountered in microfluidics, Eq. 16.22 may well be the most widely known. Please note that we have (indirectly) already derived this equation when discussing the flow rate through circular cross-section. Thus Eq. 16.8 and Eq. 16.22 are identical.

Shear Stress Profile. Next we will calculate the shear force profile. The shear stress is related to the velocity profile by Eq. 9.5 and can therefore be calculated from Eq. 16.19 as

τ=ηR24ηdpdzddr((rR)21)=12dpdzr

si39_e  (Eq. 16.23)

Compared to the velocity profile, this function is not symmetrical although it must be in order to yield physically reasonable results. The calculation is correct as the radius cannot have negative values in reality. However, when plotting the profiles, we may want to plot the parabola shape symmetrically in which case we insert negative values for the radius. In listing 16.2, line 18 we let Maple find this derivative as tau and define again a helper function tau2 in line 21. This function ensures that the shear force is positive using the function abs, which returns the absolute value of a function. Furthermore tau2 also ensures that the output of the function is in the correct units, i.e., in N m−2. Starting from line 24 we plot this function using the two tube diameters used earlier. These plots are shown in Fig. 16.8.

f16-08-9781455731411
Fig. 16.8 Shear stress profiles calculated for two tubes with different diameters. The fluid used was water, and the pressure drop along the z-axis is given.

16.3.3 Derivation Using a Force Balance Approach

The second potential approach to finding the Hagen-Poiseuille flow profile is by applying a force balance to a volume segment of the fluid. As we will see, this method is appealing because we do not have to solve a second-order but only a first-order ODE, which is very easy to solve.

Derivation. The derivation follows the schematic shown in Fig. 16.9. The forces acting on a cylindrical infinitesimal volume of length dx and radius dr are the pressure forces acting on the circular surfaces (the difference in which originates from the pressure drop) and the shear stress forces that represent the fluid friction acting on the surface of the volume. The force balance is given by

f16-09-9781455731411
Fig. 16.9 Schematic of the derivation of the Hagen-Poiseuille profile using a force balance approach.

(p1p2)AfrontτdAcylinder!__0

si40_e  (Eq. 16.24)

where

p2=p1+pzdz

si41_e  (Eq. 16.25)

Afront=πr2

si42_e  (Eq. 16.26)

dAcylinder=2πrdz

si43_e  (Eq. 16.27)

τ=ηvr

si44_e  (Eq. 16.28)

where we have used Eq. 9.5 to convert the shear stress into a function of the velocity and thus restrict the approach to Newtonian fluids. Inserting Eq. 16.25, Eq. 16.26, Eq. 16.27, and Eq. 16.28 into Eq. 16.24 yields

pzπr2dzηvr2πrdz!__0pzr=2ηvr

si45_e  (Eq. 16.29)

Analytical Solution. Eq. 16.29 is in fact not a PDE but a simple first-order ODE that can be solved by separation of variables to result in

2ηdv=dpdzrdr2ηv=dpdz12r2+c

si46_e  (Eq. 16.30)

where we can derive the integration constant from the no-slip condition at the wall, i.e., v (R) = 0 from which we find

v(r)=14ηdpdz(r2R2)=R24ηdpdz((rR)21)

si47_e  (Eq. 16.31)

As we can see Eq. 16.31 and Eq. 16.20 are identical. We have thereby succeeded in deriving the Hagen-Poiseuille flow profile using this alternative approach.

16.3.4 Average Flow Velocity

As we have seen, the Hagen-Poiseuille flow profile takes the form of a parabola (see Eq. 16.20) with the maximum flow rate in the center of the tube. In many applications it may be advantageous not to work with this velocity profile but with an average velocity vav that could, e.g., be simply multiplied with the cross-section area A = πR2 to obtain the flowrate Q. This average velocity is encountered fairly often because it makes the calculation significantly simpler. We will therefore shortly derive this velocity. The simplest derivation stems from the fact that the flowrate through the cross-section must be identical:

Qav=πR2vav

si48_e  (Eq. 16.32)

Qeffect.=0Rv(r)2πrdr

si49_e  (Eq. 16.33)

where we have used the infinitesimal area of the circle (see Eq. 3.104) for calculating the effective flowrate using the velocity profile. Please note that this is the approach we also took when deriving the flowrate in section 16.3.2. Obviously Eq. 16.32 and Eq. 16.33 must be equal in which case we find

vav=2R20Rv(r)rdr=12ηdpdx0Rr3R2rdr=12ηdpdx[r44R2r22]0R=12ηdpdx(R24R22)=R28ηdpdx

si50_e  (Eq. 16.34)

where we note that

vav=vmax2

si51_e  (Eq. 16.35)

which allows us to write the velocity profile also as

v(r)=2vav((rR)21)=2vav(1(rR)2)

si52_e  (Eq. 16.36)

We will use this notation on several occasions as it is slightly more compact.

16.3.5 Darcy Friction Factor

Darcy Friction Factor for Laminar Hagen-Poiseuille Flow. You may occasionally come across a notation of Eq. 16.34 that can be derived by rearrangement and is given as

dpdz=vav8ηR2ReRe=vav32ηd22Re2Re=vav64ηd2Re2Re=vavfη2R2Re

si53_e  (Eq. 16.37)

where we have introduced a factor that is referred to as the Darcy friction factor f for Hagen-Poiseuille flow. It is given by

fHagen-Poiseuille=64Re

si54_e  (Eq. 16.38)

The Darcy friction factors are experimentally determined coefficients that allow a correlation between average flow velocity (and therefore volume flowrate) and pressure drop expressed solely by constants and the Reynolds number. This you can see from Eq. 16.37. Eq. 16.38 is the friction factor for laminar Hagen-Poiseuille flow. It is only valid for Reynolds numbers in laminar flow. If the Reynolds number increases to the transition regime or even to turbulent flow Eq. 16.37 can still be applied, but a more complicated equation for the Darcy friction factor would need to be found. As stated, these are usually derived empirically from experimental data. By using only complicated equations for this one factor the overall equations remain very simple. Eq. 16.37 is generally referred to as the Darcy-Weisbach1,2equation.

Higher Reynolds Numbers: Colebrook-White Equation. If Reynolds numbers are above 4000 different approximations for the Darcy friction factor have to be found. Most common is the following implicit formula that was first given by Colebrook1[5] based on earlier work by him and White [6], which is why it is commonly referred to as the Colebrook-White equation:

1f=2log10(ε3.7dH+2.51Ref)

si55_e  (Eq. 16.39)

which uses the roughness height ε and the hydraulic diameter (see section 16.3.6). Eq. 16.39 is an implicit formula, meaning that it needs to be solved numerically as there is no analytical solution for f. However, this equation can be approximated conveniently using algebra tools or even numerical calculators. However, several attempts to finding explicit analytical approximations to this equation for certain fluid mechanical cases have been made, which resulted in a wide range of approximations for f from Eq. 16.39. This work is still ongoing with recent contributions [7].

16.3.6 Hydraulic Diameter

A commonly used method for more complicated geometries is the usage of the hydraulic diameter. This virtual diameter is defined as

dH=4Alc=4cross-section areacircumference

si56_e

It essentially “converts” an arbitrary cross-section to a diameter of an (fluid mechanically more-or-less) equivalent circular cross-section. Usually, the equations derived for the Hagen-Poiseuille flow are then simply applied. However, for most profiles, the errors one makes when using the hydraulic diameter instead of the correct equations for the respective cross-sections are quite severe.

16.4 Flows in Rectangular Cross-Sections: Solution to Poisson and Laplace Equations

16.4.1 Introduction

Until now, we have considered mainly cross-sections that can be reduced to one-dimensional problems with the exception of the elliptical cross-section, where we have used a trial-and-error approach that proved to be successful. Now we will turn to the case of a rectangular channel cross-section. As we will see, this cross-section is significantly more complicated. Again, we begin with Eq. 15.14 from which we find

dpdx=η(2vxz2+2vxy2)+kx

si57_e

In the first step, we will neglect volume forces which reduces the equation to

2vxz2+2vxy2=1ηdpdx

si58_e  (Eq. 16.40)

The right-hand side of Eq. 16.40 is a constant. Before attempting to solve this equation, we will spend a moment to consider what we actually learn from it. It tells us that for any point in the cross-section the change of tangents to the velocity profile in the direction of the y-axis plus the change of the tangents to the velocity profile in the direction of the z-axis will always sum up to a constant that is dependent on the viscosity and the pressure drop driving the flow. Let us refer to these “change of the tangents” as the “change in steepness of the velocity profile”. The lower the viscosity and/or the higher the pressure drop, the bigger this sum of changes will become. In other words, with decreasing viscosity and increasing pressure drop, the flow profile will change steepness faster, i.e., the relative changes in the steepness of the velocity profiles will become bigger.

Obviously these observations do not help us in finding a solution directly. However, understanding what the equation tells us makes it somewhat more intuitive to understand the problem. We need to find an equation for which, at any given point (y, z) in the cross-section the sum of the changes along the y-axis and z-axis are the given constant. Therefore, if we expect a large change along the y-axis, the change along the z-axis must be small and v.v. We have already calculated the flow profile in an infinitesimally elongated channel to be of parabolic shape (see section 16.2). In this case the change along the y-axis was neglected.

16.4.2 Naive Approach

We can approach this problem naively and assume that Maple may be able to solve this problem for us. The following line will ask Maple to solve Eq. 16.40 for us using the command pdsolve, which is used to solve PDEs:

equ1:=diff(diff(u[x](y,z),y),y)+diff(diff(u[x](y,z),z),z) = -1/eta*dpdx;
pdsolve(equ1,u[x](y,z));

The output we will obtain will most likely look somewhat like this:

u[x](y, z) = _F1(z - I y) + _F2(z + I y) - (dxdy*y^2)/(eta)

What Maple is telling is that it expects the solution to consist of two functions that take complex argument expressions involving y and z. It also expects this function to contain a term only containing y, which mainly takes care of the constant. Obviously, this solution is not very helpful. Besides being rather vague, this solution also does not account for our boundary conditions. All together, this was not a very successful approach.

16.4.3 General Approach: Substitution and Separation of Variables

We have derived a number of solutions to commonly encountered (partial) differential equations in section 8.2 and section 8.3. We will now apply these methods for finding the solution to Eq. 16.40. In the more general terms, we will solve the Poisson equation (see section 8.1.5), which is (in general) given by Eq. 8.17 as

Δf(y,z)=g(y,z)

si59_e  (Eq. 16.41)

The Poisson equation is an inhomogeneous partial differential equation. We therefore first default to solving the homogeneous case, i.e., the Laplace equation (see section 8.1.6) given by Eq. 8.18 as

Δf(y,z)=0

si59_e  (Eq. 16.42)

We will solve this function in a two-dimensional case on a rectangular boundary 0 ≤ y ≤ w and 0 ≤ z ≤ h with the boundary conditions f (0, z) = f (w, z) = f (y, 0) = f (y, h) = 0. In fluid mechanical terms, this solution will give us the Poiseuille flow profile in channels with rectangular cross-sections of height h and width w. We assume no-slip condition which gives rise to our boundary conditions.

We proceed using a separation of variables approach using f (y, z) = Y (y) Z (z) as detailed in section 8.3.3.5

f(y,z)=Y(y)Z(z)Δf(y,z)=2Yy2Y+2Zz2Z=0

si61_e  (Eq. 16.43)

There are two main ways of approaching Eq. 16.43, both of which you are likely to encounter in fluid mechanics textbooks. It is a question of personal favor which one to choose. We will briefly outline both.

16.4.4 Single λ Approach

The first method of going about it could be called a “single λ approach”. In this case we rewrite Eq. 16.43 to

2Yy2Z=2Zz2Y1Y2Yy2=1Z2Zz2=λ2

si62_e  (Eq. 16.44)

where we again state that in order for this equation to be fulfilled, both derivatives must equal a constant, which we term −λ2. We have taken this approach in section 8.3.3.5 while solving the one-dimensional heat equation. We use λ as a squared term here in order to be exact about the signs.

Coordinate System and Boundary Conditions. Depending on the textbook you may find different versions of the solution to Eq. 16.42. Fig. 16.10 lists the most commonly encountered coordinate systems used.

f16-10-9781455731411
Fig. 16.10 Coordinate systems for deriving the Poiseuille flow profile in a rectangular channel cross-section.

Solution on the Boundsw2y+w2si63_e and 0hsi64_e. This case is depicted by the coordinate system in Fig. 16.10a. We use the following Dirichlet boundary conditions:

f(w2,z)=0Y(w2)=0f(+w2,z)=0Y(w2)=0f(y,0)=0Z(0)=0f(y,h)=0Z(h)=0

si65_e

We have already done most of the work in finding the solution to this case. Referring to Tab. 8.2 we find the following eigenvalue functions as a solution for Z (Z) with λ < 0:

Z(z)=n=1cnsin(nπzn)

si66_e  (Eq. 16.45)

which also gives us

λ=nπh

si67_e  (Eq. 16.46)

The case for Y (y) is less straightforward. Referring to Tab. 8.2 we find for λ > 0 the solution

Y (y) = c0 sinh (λ y) + c1 cosh (λ y)

However, after applying the boundary condition, we are only left with the trivial solution. This case happens surprisingly often when solving differential equations and it is due to one simple fact: We know too much of the solution. Remember, the boundary conditions are used to determine λ. We already applied two boundary conditions and solved for λ. If we apply two boundary conditions again, we will end up with a situation similar to having more equations than unknowns to solve for. In applying more boundary conditions than we actually need, we are left with only the trivial solution as a result.

However, we can also apply a different type of boundary condition. We know that the system we are investigating is symmetric along the z-axis. Therefore, it is reasonable to assume that the solution function will also be symmetric along the z-axis and thus the resulting function will have its maximum at y = 0. In this case, we can assume the first derivative of Y (y) to be zero at the origin: dYdy|y=w0=0si68_e. This is a Neumann boundary condition. Again referring to Tab. 8.2 we find the corresponding solution to be

Y(y)=c1cosh(λy)=c1cosh(nπyh)

si69_e  (Eq. 16.47)

where we have already inserted the solution for λ from Eq. 16.46. Now, Y (y) does not yet fulfill our boundary conditions Y(w2)=Y(+w2)=0si70_e. Unfortunately plugging in these boundary conditions will leave us only with the trivial solution for c1 = 0, and we need to do a little more thinking. If we recall the shape of the hyperbolic cosine function (see Fig. 3.4b) we know that this function is never 0. In order to satisfy our boundary conditions, we need to use the negative hyperbolic sine function and shift it such that it will be zero at Y(w2)andY(+w2)si71_e. As the function is symmetric, we only need to account for one of the points, which makes life a little easier. We assume our function to look like the following:

Y(y)=c2c1cosh(nπyh)Y(w2)=c2c1cosh(nπw2h)!__0c2=c1cosh(nπw2h)Y(y)=c1(cosh(nπw2h)cosh(nπyh))=c1cosh(nπw2h)(1cosh(nπyh)cosh(nπw2h))=c3(1cosh(nπyh)cosh(nπw2h))

si72_e  (Eq. 16.48)

where the function now satisfies the boundary condition. Having derived Y (y) and Z (z) we can now assemble the homogeneous solution to Eq. 16.42:

fh(y,z)=Y(y)Z(z)=c1c3(1cosh(nπyh)cosh(nπw2h))n=1cnsin(nπzh)=n=1cn(1cosh(nπyh)cosh(nπw2h))sin(nπzh)

si73_e  (Eq. 16.49)

This is the solution to the homogeneous Poisson (i.e., Laplace equation) given by Eq. 16.42. Obviously, the constants cn still need to be determined. We will do this by taking into account the term right-hand side of Eq. 16.41, i.e., g (y, z). This will provide us with the inhomogeneous solution we are seeking.

Inhomogeneous Solution to the Navier-Stokes Equation. We first start by forming the second-order partial differentials of Eq. 16.49 that we need in order to take into account the right-hand side equation g (y, z). In our case we find

2fhy2=n=1cn(nπh)2cosh(nπyh)cosh(nπw2h)sin(nπzh)2fhz2=n=1cn(1cosh(nπyh)cosh(nπw2h))+(nπh)2sin(nπzh)2fhy2+2fhz2=n=1cn(nπh)2sin(nπzh)(cosh(nπyh)cosh(nπw2h)+1cosh(nπyh)cosh(nπw2h))=n=1cn(nπh)2sin(nπzh)

si74_e  (Eq. 16.50)

From Eq. 16.41 we know that this must equal the right-hand side g (y, z). If you look carefully at Eq. 16.50 you will see that the sine series that we already had in our initial result (Eq. 16.49) is still there. It therefore makes perfect sense to expand the right-hand side of Eq. 16.41 to a sine series as described in section 4.3.3.8. We can then determine cn simply by comparing the coefficients. Referring to Eq. 16.40 you see that the right-hand side is not a function of y and z:

g(y,z)=1ηdpdx=const.

si75_e

We therefore merely need to Fourier expand this constant to a sine function as detailed in Eq. 4.34. The expansion should be along the z-axis since this is the axis along which our sine function was expanded (see Eq. 16.49). In this case we find

1ηdpdx=n=04ηπdpdx1(2n+1)sin((2n+1)πzh)

si76_e  (Eq. 16.51)

Now we need to match the coefficients of Eq. 16.50 and Eq. 16.51. First we find that we only need to take into account all terms for odd values of n since the even terms vanish when expanding a constant to a Fourier series (see section 4.3.3.3). We therefore find

n=04ηπdpdx1(2n+1)sin((2n+1)πzh)!__n=0cn((2n+1)πh)2sin((2n+1)πzh)

si77_e

from which we find

4ηπdpdx1(2n+1)!__cn((2n+1)πh)2cn=4h2ηπ3dpdx1(2n+1)3

si78_e  (Eq. 16.52)

which gives us the missing constants cn. We are now able to formulate the overall solution to the Poisson equation in rectangular channel cross-sections given by Eq. 16.49 and Eq. 16.52

f(y,z)=n=04h2ηπ3dpdx1(2n+1)3(1cosh((2n+1)πyh)cosh((2n+1)πw2h))sin((2n+1)πzh)=4h2ηπ3dpdxn=01(2n+1)3(1cosh((2n+1)πyh)cosh((2n+1)πw2h))sin((2n+1)πzh)

si79_e  (Eq. 16.53)

Solutions on the Boundsw2y+w2andh2z+h2si80_e. You may come across different solutions for the Poiseuille flow profile in rectangular cross-sections if a different coordinate system is used. The case we will now discuss is depicted in Fig. 16.10b. We use the following Dirichlet boundary conditions:

f(w2,z)=0Y(w2)=0f(+w2,z)=0Y(w2)=0f(y,h2)=0Z(h2)=0f(y,+h2)=0Z(+h2)=0

si81_e

We will start by finding the solution to Z (z), which we can simply look up in Tab. 8.2 after applying the boundary conditions:

Z(z)=n=0cncos((2n+1)πhz)λ=(2n+1)πh

si82_e

Now we can use the same derivation for Y (y) we took for the coordinate system of Fig. 16.10a. Again we find

Y(y)=c2c1cosh((2n+1)πyh)Y(w2)=c2c1cosh((2n+1)πw2h)!__0c2=c1cosh((2n+1)πw2h)Y(y)=c1(cosh(2n+1)πw2h)cosh((2n+1)πyh)=c1cosh((2n+1)πw2h)(1cosh((2n+1)πyh)cosh((2n+1)πw2h))=c3(1cosh((2n+1)πyh)cosh((2n+1)πw2h))

si83_e  (Eq. 16.54)

which is almost the same as Eq. 16.48 but uses a different λ. Therefore we find as the homogeneous solution:

fh(y,z)=Y(y)Z(z)=c1c3(1cosh((2n+1)πyh)cosh((2n+1)πw2h))n=0cncos((2n+1)πzh)=n=0cn(1cosh((2n+1)πyh)cosh((2n+1)πw2h))cos((2n+1)πzh)

si84_e  (Eq. 16.55)

Obviously, we now have a cosine series, which is why we expand the right-hand side of Eq. 16.40 to a cosine Fourier series according to Eq. 4.42, in which case we find

1ηdpdx=n=0(1)n4(2n+1)π4ηπdpdxcos((2n+1)πzh)

si85_e  (Eq. 16.56)

By comparison with Eq. 16.55 we can now determine the missing coefficient cn. However, we must first calculate the left-hand side of Eq. 16.55 for which we need the partial derivatives of Eq. 16.55:

2fhy2=n=0cn((2n+1)πh)2cosh((2n+1)πyh)cosh((2n+1)πw2h)cos((2+1)πzh)2fhz2=n=0cn(1cosh((2n+1)πyh)cosh((2n+1)πw2h))+((2n+1)πh)2cos((2n+1)πzh)2fhy2+2fhz2=n=0cn((2n+1)πh)2cos((2n+1)πzh)(cosh((2n+1)πyh)cosh((2n+1)πw2h)+1cosh((2n+1)πyh)cosh((2n+1)πw2h))=n=0cn((2n+1)πh)2cos((2n+1)πzh)

si86_e  (Eq. 16.57)

Comparing Eq. 16.56 and Eq. 16.57 we find the coefficients cn to be

n=0(1)n4(2n+1)π4ηπdpdxcos((2n+1)πzh)!__n=0cn((2n+1)πh)2cos((2n+1)πzh)

si87_e

from which we find

(1)n4(2n+1)π4ηπdpdx!__cn((2n+1)πh)2cn=(1)n4h2ηπ3(2n+1)3dpdx

si88_e

which gives us the missing constants cn. Therefore we find the solution to the Poisson equation in rectangular channel cross-sections using the coordinate system depicted in Fig. 16.10b to be

f(y,z)=n=0(1)n4h2ηπ3(2n+1)3dpdx(1cosh((2n+1)πyh)cosh((2n+1)πw2h))cos((2n+1)πzh)=4h2ηπ3dpdxn=0(1)n1(2n+1)3(1cosh((2n+1)πyh)cosh((2n+1)πw2h))cos((2n+1)πzh)

si89_e  (Eq. 16.58)

Eq. 16.58 is a second version of the solution to the Navier-Stokes equation in rectangular channel cross-sections.

16.4.5 Two λ Approach

We have now derived two solutions to Eq. 16.40 both of which employed a single λ. In both cases, we have done a single eigenvalue expansion, i.e., we obtained a single set of eigenfunctions as solutions. These solutions were once derived as a sine series and once as a cosine series.

There is a second approach to solving Eq. 16.40 using two eigenvalue expansions. In this case, we will obtain two different values, λy and λz.

Solution on the Bounds 0 ≤ y ≤ w and 0 ≤ z ≤ h. For this approach we will use the coordinate system depicted in Fig. 16.10c in which case we find the following boundary conditions:

f(0,z)=0Y(0)=0f(w,z)=0Y(w)=0f(y,0)=0Z(0)=0f(y,h)=0Z(h)=0

si90_e

Referring to Tab. 8.2 we find the following solutions for this case after applying the boundary conditions:

Y(y)=n=1cnsin(nπyw)Z(Z)=m=1cmsin(mπzh)

si91_e

In which case we find the homogeneous solution to be

fh(y,z)=Y(y)Z(z)=n=1cn(nπyw)m=1cmsin(mπzh)=n=1m=1cnmsin(nπyw)sin(mπzh)

si92_e  (Eq. 16.59)

As you can see, we have used two eigenvalue expansions that give rise to two different lambda terms and we are therefore left with two sums. The derivation is straightforward but usage of a second sum may well be considered a disadvantage.

We will now form the two second-order partial differentials for the left-hand side of Eq. 16.40, which is given by

2fhy2=n=1m=1cnm(nπw)2sin(nπyw)sin(mπzh)2fhz2=n=1m=1cnm(mπh)2sin(nπyw)sin(mπzh)2fhy2+2fhz2=n=1m=1cnm((nπw)2+(mπh)2)sin(nπyw)sin(mπzh)

si93_e  (Eq. 16.60)

Now, we must expand the right-hand side of Eq. 16.40 to a Fourier sine series in two dimensions according to Eq. 4.37

1ηdpdx=16ηπ2dpdxn=0n=01(2m+1)(2n+1)sin((2n+1)πyw)sin((2m+1)πzh)

si94_e  (Eq. 16.61)

As we can see, this function only requires coefficients for odd values of n and m. By coefficient comparison of Eq. 16.60 and Eq. 16.61 we now find the missing coefficients cnm as

n=0n=016ηπ2dpdx1(2m+1)(2n+1)sin((2n+1)πyw)sin((2m+1)πzh)!__n=0m=0cnm(((2n+1)πw)2+((2m+1)πh)2)sin((2n+1)πyw)sin((2m+1)πzh)16ηπ2dpdx1(2m+1)(2n+1)!___cnm(((2n+1)πw)2+((2m+1)πh)2)cnm=16ηπ2dpdx1(2n+1)(2m+1)(((2n+1)πw)2+((2m+1)πh)2)

si95_e

Inserting this expression into Eq. 16.59 yields the solution to the inhomogeneous Navier-Stokes equation Eq. 16.40 as

f(y,z)=16ηπ2dpdxn=0m=0sin((2n+1)πyw)sin((2m+1)πzh)(2n+1)(2m+1)(((2n+1)πw)2+((2m+1)πh)2)

si96_e  (Eq. 16.62)

16.4.6 Visualization: Single λ Approach

Before turning to calculating the flow rate we will quickly visualize the equations we just derived. We start with the equations obtained from the “one λ approach”. We will start by writing a small Maple worksheet that will output the profiles for us.

Maple Worksheet. As you can see, the listing is not very long. First, we restart the Maple server and include Core.txt (see line 1). In line 4 we define the velocity profile. This is Eq. 16.53 with the added dimensions. We assume the viscosity eta in mPa s, the pressure drop in mbar mm−1, the height h, and width w of the channel in µ m. Furthermore, we include the expansion order nmax to which the Fourier series should be expanded. The function uses simplify command in order to ensure that the output of the function is dimensionless by dividing it by mm s−1, which is the output unit. In order to simplify plotting, we also apply the evalf command in order to turn all analytical expressions to numbers for plotting.

In line 7 we plot the flow profile for different pressure drops assuming water as the liquid of choice. The channel dimensions are chosen arbitrarily as 100 µ m for w and 50 µ m for h. The output of this line is Fig. 16.11. As you can see, for higher pressure drops the velocity maximum reaches significant values. For a pressure drop of − 0.1 mbar mm−1 we achieve a center velocity of almost 2.8 mm s−1.

f16-11-9781455731411
Fig. 16.11 Flow profiles vx (y, z) calculated for water in a rectangular cross-section channel with increasing pressure drops of − 0.001 mbar mm−1 (dark blue), − 0.01 mbar mm−1 (medium green), and − 0.1 mbar mm−1 (light yellow). The channel has a width of 100 µm and a height h of 50 µm.

Normalized Flow Profiles for Different Aspect Ratios. Obviously, the equations we found for all cases are Fourier series. Thus the precision of our solutions depends on the chosen expansion order of the Fourier series. Before turning to simplifications of the equations found, we will look at the normalized flow profiles in rectangular channels with different cross-sections. For this, we expand Eq. 16.53 to nmax = 50 and plot the profiles for aspect rations r of 0.1, 1, 10, and 100. The plots are shown in Fig. 16.12. They are normalized to the maximum flow rate (on the y-axis) and to the width on the y-axis and the height on the z-axis. You will note the different aspect ratios showing as different values for z whereas y is always plotted as relative width with −0.5 ≤ y ≤ 0.5.

f16-12-9781455731411
Fig. 16.12 Normalized flow profiles in rectangular channel cross-sections with different aspect ratios r.

As you can see, for small aspect ratios (r = 0.1) as well as for large aspect ratios (r = 10, r = 100) we obtain flow profiles that resemble the flow profiles of the planar infinitesimally extended channel (see section 16.2). This is to be expected as the problem becomes pretty much one-dimensional at very small and very large aspect ratios. At aspect ratios close to 1 we can see the shape of the flow profile more clearly. These are the more challenging cases for fluid mechanical problems as it becomes virtually impossible to use one-dimensional approximations for the flow profiles at aspect ratios around 1.

Listing 16.3

[Maple] Listing for calculating flow profiles in rectangular channels. A digital version of the listing can be found under the name RectangularChannel.mw in the supplementary material.

1 restart:read "Core.txt":
2
3 #this is the flow profile
4  u[x]:=(y,z,w,h,eta,dpdx,nmax)->evalf(simplify((-4*(h*Unit(um))^2)/(eta*Unit(mPa*s)*Pi^3)*dpdx*Unit (mbar/mm)*sum((1/((2*n+1)^3)*(1 - cosh((2*n+1)*Pi*y/h)/cosh((2*n+1)*Pi*1/2*w/h))*sin((2*n+1)* Pi*z/h)/Unit(mm/s),n=0..nmax))));
5
6 #plot for different pressure drops
7 quickplot3d([(y,z)->u[x](y,z,100,50,1,-0.001,10),(y,z)->u[x](y,z,100,50,1,-0.01,10),(y,z)->u[x](y, z,100,50,1,-0.1,10)],0..50,0..50,"y [um]","z [um]",typeset(v[x]^empty,[mm/s]));

Approximations. Having studied the normalized flow profiles for different aspect ratios, the following question arises: do we really need to use the full Fourier series expansion if calculating the flow profile? Obviously, the answer to this question depends on how exact the solution needs to be. In this section, we will briefly study how close simplified versions of Eq. 16.53 can be if only expanded to small expansion orders. We will study the cases for nmax = 2 and nmax = 5, comparing them with the close-to-exact solution for nmax = 10. We will do this for all four aspect ratios used in Fig. 16.12.

In order to get a better view on the profiles, we only plot −0.5 ≤ y ≤ 0 so only half of the profile. This allows us to “look into” the profiles and see how big our mistakes due to simplification. The plots are shown in Fig. 16.13. As you can see, both approximations are surprisingly exact for aspect ratios of around and below 1. However, they become less and less exact as the aspect ratios increase. This is due to the fact that we did the Fourier sine expansion along the z-axis during derivation. Therefore, in order to give an exact solution, a decent expansion order is required in order to correctly represent the profile along the z-axis. On the other hand, if the elongation of the channel along the y-axis dominates, there is no reason to add a significant number of terms as the hyperbolic cosine functions dominate the flow profiles.

f16-13-9781455731411
Fig. 16.13 Approximations for the flow profiles in rectangular channel cross-sections with different aspect ratios r. Displayed are the profiles for an expansion order of 2 (dark blue), 5 (medium green), and the close-to-exact solution for 50 (light yellow). As can be seen, the approximations are nearly exact for shallow channels and become more and more inexact the higher the aspect ratio becomes.

Fig. 16.14 plots the relative error vx,nmax={1,2}vx,nmax=50vx,nmax=50,maxsi97_e we make when simplifying the velocity profiles by taking an expansion order to nmax = 2 and nmax = 1. As you can see, for nmax = 1 the error is in the range of 1.6 % for channels with aspect ratios of 0.1 and about 3 % for channels with aspect ratios of 1. When using nmax = 2, the error becomes smaller, amounting to only around 0.8 % for channels with aspect ratios of 0.1 and to about 1.3 % for channels with aspect ratios of 1. For the stated reasons, it is not advisable to use these simplifications for channels with higher aspect ratios. Also, if the velocity profile will be used for additional calculations (e.g., the calculation of the flow rate) it is also not advisable to use these simplifications since, obviously, the small (local) derivation from the exact solution will quickly add up to significant overall errors.

f16-14-9781455731411
Fig. 16.14 Simplified normalized velocity profiles for channels with rectangular cross-sections for aspect ratios r ≤ 1. The expansion order for this simplification is nmax = 1 and nmax = 2. The displayed error is calculated asvx,nmax={1,2}vx,nmax=50vx,nmax=50,maxsi1_e

However, if a simplification is required and the aspect ratio of the channel is in the range of r ≤ 1 these very simple approximations for the velocity profile can be used:

fnmax=1(y,z)=4h2ηπ3dpdx((1cosh(πyh)cosh(πw2h))sin(πzh)+127(1cosh(3πyh)3cosh(πw2h))sin(3πzh))fnmax=2(y,z)=4h2ηπ3dpdx((1cosh(πyh)cosh(πw2h))sin(πzh)+127(1cosh(3πyh)3cosh(πw2h))sin(3πzh)+1125(1cosh(5πyh)5cosh(πw2h))sin(5πzh))

si98_e

For channels with aspect ratios bigger than 1 similar approximations can be obtained by performing the eigenvalue expansion along the y-axis instead of along the z-axis, as we have done in section 16.4.4. This will give an expression very similar to Eq. 16.53 but with the hyperbolic cosine as function along the z-axis. Then similar approximations to the ones we have just derived can be found.

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

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