You are on page 1of 19

2nd Year Fluid Mechanics, Faculty of Engineering and Computing, Curtin University

FLUID MECHANICS 230


For Second-Year Chemical, Civil and Mechanical Engineering FLUID MECHANICS LECTURE NOTES CHAPTER 6 VISCOUS FLOW IN PIPES
6.1 Introduction In this chapter, we will consider viscous incompressible flow in pipes where the fluid is confined and bounded. Pipe systems are widely used in practice. Typical examples include drinking water distribution pipe systems, oil pipe lines etc. Figure 6-1 presents the schematic diagram of a typical pipe system.

Figure 6-1 A typical pipe system [1] As shown in Figure 6-1, a pipe system may include 1. Individual straight pipes 2. Pipe connectors (such as Tee-union, elbow connector etc.), for connecting pipes 3. Flow rate control devices (such as valves) for adjusting the flow rate 4. Inlet and outlet 5. Pumps which add energy into the fluid where items 2, 3 and 4 are often called pipe components. 6.2 Real pipe flow For inviscid, incompressible, steady and irrotational flows, Bernoullis Equation applies V 2 P + gz + = const , along a streamline which can also be written as 2 V 2 ( P + gz + )=0. (E6-1) 2 Equation E6-1 indicates that for ideal fluid ( = 0 ), if there is no change in the sum of the fluid elevation and dynamic pressure, i.e. ( gz ) + ( V 2 2) = 0 , the overall pressure drop across a pipe system should be zero.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 1 of 19

However, for real fluid, energy loss hence pressure drop PL across a pipe system is inevitable because of the shear force (friction) between the pipe and the fluid. This friction is a result of the viscous nature of real fluid ( 0 ). To deliver the fluid through a pipe system, pumps are often introduced into the pipe system in order to provide pressure rise Pwork . Therefore, E6-1 is no longer applicable to real pipe flows and we need to consider both PL and Pwork , i.e.

) + Pwork PL = 0 . (E6-2) 2 The pressure drop PL can be due to pressure drop in straight pipes (Item 1 in Figure 6-1), called major loss, and pressure drop in pipe components (Items 2, 3 and 4 in Figure 6-1), called minor loss. Therefore, we have (E6-3) PL = PL ,major + PL ,min or . Equation E6-2 shows that to deliver fluid through a pipe system, a pump needs to be properly selected to provide enough pressure rise to overcome the pressure drop due to friction in the pipe system, the changes in elevation and the dynamic pressure. It indicates that to design and operate a real pipe system, we need to have sufficient knowledge on: 1. the relationship among pressure drop across a straight pipe (i.e. major loss), the pipe properties and flow properties (This is discussed in Chapter 6) 2. the relationship among pressure drop across pipe components (i.e. minor loss), the properties of the pipe component and flow properties (This is discussed in Chapter 7) 3. the energy gain by devices such as pumps (This is the topic of Chapter 10)
Head forms of E6-1, E6-2 and E6-3

( P + gz +

V 2

These equations can also be expressed in head forms as the following: V2 P ( +z+ )=0 g 2g
( V2 P +z+ ) + hwork hL = 0 g 2g hL = hL ,major + hL ,min or .

(E6-1a) (E6-2a) (E6-3a)

In the head form, each of the terms has the units of length and represents a certain type of head. For example, the elevation, z, is related to the potential energy and is called elevation head. The pressure term, P / g , is called static (or pressure) head and represents the height of a column of the fluid that is needed to produce the pressure, P. The velocity term, V 2 / 2 g , is the dynamic (or velocity) head and represents the vertical distance needed for the fluid to fall freely if it is to reach velocity V from rest. Similarly, hwork and hL are termed as pump head and head loss, respectively.
6.3 Pioneer work on measuring pressure drop across a pipe 6.3.1 Pressure-drop test [2] Figure 6-2 illustrates is a simple experimental setup for measuring pressure drop across a pipe. Liquid flew from the tank (by elevation energy) to the pipe. There would be a long section where the flow was not uniform, before the fluid entering the test section to produce a uniform flow. The test section has a length of x and the pressures at both ends were measured as P1 and P2.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 2 of 19

A flow-regulating valve was introduced to control the flow rate so that tests could be conducted to find the correlation between the pressure gradient across the test section and the fluid flow rate in pipe. Figure 6-3 presents the original experimental results of Osborne Reynolds in 1883 of one specific fluid and one specific pipe [3]. Extensive experiments showed that these observations are generic for pipe flow, regardless of the type of liquid and kind of pipe used in such experiments.

Figure 6-2 Experimental setup for pressure-drop test [2].

Figure 6-3 Measured pressure gradient

P P P (i.e. 2 1 ) of a specific pipe as a function of x x volumetric flow rate Q of a specific fluid [2], originally from reference [3].
Chapter 6 Page 3 of 19

Prepared by Dr Hongwei Wu, since 2006

Depending on the flow rate Q, Figure 6-3 shows that the measured pressure gradient of the test section may fall into one of the three regions: Region I: At a low flow rate, the pressure gradient is proportional to flow rate; Region II: At an intermediate flow rate, the flow behaviour seemed to be unpredictable so that the experimental data are scattered within the regions of the two curves which are extrapolated from those in Region I and III. Region III: At a high flow rate, the pressure gradient is proportional to the flow rate to the 1.8 (for smooth pipe) or 2.0 power (for rough pipe); The questions are: why are there three regions in Figure 6-3? How should we interpret Figure 6-3 and obtain enough knowledge on pipe flows to guide the design of pipe systems?
6.3.2 Laminar, transitional and turbulent flow in pipe [1] In 1883, Osborne Reynolds [3] did the pioneer work to understand Figure 6-3. Flow in the three regions were visualised using a transparent pipe with a dye streak as a tracer. The flow patterns of the three regions in Figure 6-3 are shown in Figure 6-4 while the time dependence of fluid velocity at point A is shown in Figure 6-5. 1. Laminar flow: the dye streak remains a steady line as it flows through the pipe. All the flow motion is in axial direction, there is no mixing perpendicular to the axis of the pipe. This is observed at low flow rates, i.e. Region I in Figure 6-3. 2. Transitional flow: the dye streak fluctuates in time and space, and intermittent bursts of irregular behaviour appear along the streak. The flow motion is not solely in axial direction anymore. Some mixing perpendicular to the axis of the pipe may occur. This is observed at higher flow rates, i.e. Region II in Figure 6-3. 3. Turbulent flow: the dye streak spreads across the entire pipe in a random fashion. The flow motion is chaotic in all directions, causing rapid, crosswise mixing. This is observed at high enough flow rates, i.e., Region III in Figure 6-3. VD It has been found that the Reynolds number, Re = , is a key dimensionless number that

determines the flow in pipe as laminar, transitional or turbulent flow. Reynolds number indicates the ratio between inertial force and viscous force exert on the fluid in pipe. At low Reynolds numbers, viscous force dominates and the flow motion is well-defined. The flow is laminar flow. At high Reynolds numbers, inertial force dominates so that flow becomes turbulent and chaotic. The flow becomes turbulent flow. With the intermittent Reynolds numbers, the flow is in transition from laminar to turbulent, called transitional flow.

Figure 6-4 Laminar, transitional and turbulent flows in pipe. Figure is from Reference [1], the work was originally done by Osborne Reynolds in 1883.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 4 of 19

Figure 6-5 Time dependence of fluid velocity at point A for laminar, transitional and turbulent flows in a pipe; Figure is from reference [1]. It should be pointed out that besides the flow rate (corresponding to the average velocity V), the flow character is also determined by fluid density , viscosity and the pipe diameter D (i.e., the fluids Reynolds number). For flow in a round pipe, we have Re < 2100 Laminar flow 2100 < Re < 4000 Transitional flow Re > 4000 Turbulent flow Therefore, the existence of the three flow regimes leads to the experimental observations of the three regions in Figure 6-3. The first region corresponds to laminar pipe flow (Re < 2100). The second region, where pressure drop is unpredictable, is due to transitional pipe flow (2100 <Re < 4000). The third region corresponds to the turbulent pipe flow (Re > 4000). We will discuss these three regions further in Section 6.7 of this chapter.
6.3.3 Entrance region flow and fully developed pipe flow When fluid enters a pipe, there exists a section, where the pipe flow is not fully developed, i.e., a uniform flow can only be produced after this section. The flow in this section is called entrance region flow. This is why during the pressure drop test in Figure 6-2, the test was done across the section where a uniform flow was produced. Such a uniform flow is called fully developed pipe flow.

The length of the entrance flow region of a pipe flow, Le, is generally short and a function of Reynolds number. Le for laminar flow (E6-4a) = 0.06 Re D Le = 4.4(Re)1 / 6 for turbulent flow (E6-4b) D The pressure drop in the entrance region is higher than the fully developed region due to extra energy loss when fluid flows into the entrance. It is generally measured by experiments and will be discussed in Chapter 7. This chapter focuses on fully developed pipe flows.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 5 of 19

6.4 Fully developed laminar pipe flow At very low flow rates (fluid velocities), viscous effects dominate, the flow is laminar flow. In the fully develop region, the viscous forces are in equilibrium with pressure forces so that the velocity profile and pressure gradient remain constant along the pipe.

6.4.1 Force balance It is simple to do force analysis as conditions in the fully developed laminar pipe flow are constant. Figure 6-6a illustrates the force balance system. For a fluid element at any given time, both pressure forces and shear force in action. Since there is no acceleration, these forces in the x-direction must sum to zero. At a radius r, we have FP = P(r2) (P+P)( r2) Pressure force: Fs = (2rx)(shear stress at r) Shear force:

Because there is no acceleration, according to Newtons 2nd motion law F = ma, F = 0, i.e., the two forces are balanced. We can equate the two forces and take the limit x 0 , yielding r dP Shear stress at r, (r): (r ) = (E6-5) 2 dx
It should be noted that since the force balance is generic, E6-5 is applicable to all types of fluids, both Newtonian and non-Newtonian flows. Equation 6-5 indicates that 1. The pressure gradient across a pipe is a constant. 2. As shown in Figure 6-6b, the shear stress at the pipe centreline (r = 0) is zero and the shear stress reaches maximum at the pipe wall (r = R). The shear stress is linear between the centreline and the wall; R dP . The wall stress is proportional to pressure 3. The wall shear stress is w = 2 dx gradient across the pipe.

(r)

(a) (b) Figure 6-6 (a) Force balance system in pipe flow; (b) Distribution of shear stress in pipe 6.4.2 Velocity profile For Newtonian flow, assume at a radius r, the flow velocity is V(r), Newtons law of viscosity applies. We can write the shear stress as dV (r ) (r ) = . Newtons law of viscosity: dr Substituting into E6-5, we have dV (r ) r dP = ( ) dr 2 dx
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 6 of 19

where

dP is a constant. Therefore, we can integrate the equation and apply the boundary dx conditions V(R) = 0, producing the following velocity profile: 2 R 2 dp r V (r ) = ( ) 1 (E6-6) 4 dx R

Equation E6-6 indicates that 1. The velocity at the pipe wall is zero, i.e., the fluid is non-slip at the wall; 2. The fluid velocity Vc along the pipe centreline (where r = 0), Vc, reaches maximum. R 2 dp ( ) ; We have: Vc = Vmax = 4 dx 3. The velocity profile is parabolic, as it can be expressed in terms of Vc, r 2 V (r ) = Vc 1 , as shown in Figure 6-7. R

Figure 6-7 Velocity profile of laminar pipe flow

Figure 6-8 Flow through a pipe

6.4.3 Flow rate The flow is axisymmetric about the pipe centreline. As shown in Figure 6-8, through a ring at radius r when its thickness dr is thin enough at the cross-section (with an area of dA = 2rdr ) of a pipe, the velocity can be considered as constant. Therefore, the flow rate Q through the pipe can be calculated as
Q=

pipe

V (r )dA
R 0 2 R 2 dp r ( ) 1 2rdr 4 dx R R

r 2 = 2Vc 1 rdr 0 R

i.e. (D is the diameter of the pipe, D = 2R)

Q=

R 2Vc
2

R 4 dp D 4 dp = 8 dx 128 dx
Chapter 6 Page 7 of 19

Prepared by Dr Hongwei Wu, since 2006

The average velocity V can be calculated as R 4 dp Q Vc 8 dx R 2 dp D 2 dp V = = = = = 8 dx 32 dx A 2 R 2 If the pressure drop across a pipe with a length of L is P, i.e., dp P , we have = dx L (E6-7)

Q=

D 4 P 128 L

D 2 P 32 L Note, E6-7 and E6-8 are ONLY suitable for laminar flow! V =

(E6-8)

6.5 Fully developed turbulent pipe flow 6.5.1 Randomness of turbulent pipe flow In turbulent flow (Re > 4000), the fluid experiences random, chaotic motion, including strong eddy transport on a macro scale, compared with the molecular motion in laminar flow. As a result, there is much more dissipation of energy as fluid molecules experience velocity changes in all directions. We can consider the two velocity components, i.e. V = ui + vj . Figure 6-9 shows an example of the axial component, u(t), measured at a given location.

Figure 6-9 Axial component of velocity at a given location in turbulent pipe flow Although the flow is chaotic, the velocity can be described in terms of a mean value (denoted with an overbar) on which the fluctuations (denoted with a prime) are superimposed. u = u + u' (E6-9a) where 1 t 0 +T u = u dt T t0 1 t0 +T u ' = u 'dt = 0 . T t0

Prepared by Dr Hongwei Wu, since 2006

Chapter 6 Page 8 of 19

Similarly, in the radial direction of the pipe flow, we have v = v + v' where v=0 1 t0 +T v ' = v'dt = 0 . T t0

(E6-9b)

6.5.2 Turbulent shear stress In turbulent flow, in addition to the motion of fluid particles discussed in laminar flow, there is also motion across the flow direction. The parcels of fluid experience relatively larger shear stress, resulting in momentum transfer. Figure 6-10 illustrates the momentum transport in a control volume in turbulent pipe flow.

Figure 6-10 Momentum transport in turbulent pipe flow In the x-direction we have:
Sum of applid Rate of momentum increase Net rate of momentum + forces + transfered int o the CV = 0 in the CV

1. We need to consider: Pressure force: FP = P(r2) (P+P)( r2) Viscous shear force: Fs = ( 2rx) Sum of applied forces: 2rx r2P 2. Rate of momentum increase in the CV: 0 3. Net rate of momentum transferred into the CV: v ' (u + u ' )2rx
Therefore, we get 2rx r 2P + v ' (u + u ' )2rx = 0 Collecting the terms, rearranging then taking the limit x 0 , dP r v ' u u ' v ' = dx 2 Applying time-averaging gives dP r v ' u u 'v ' = dx 2
As v ' = 0 , it reduces to dP r + ( u ' v ' ) = dx 2
Prepared by Dr Hongwei Wu, since 2006

(E6-10)
Chapter 6 Page 9 of 19

du . From E6-5, if we consider a total shear stress at r, total , which include fluid dr viscous shear stress and the shear stress due to eddy transport, we have dP r (E6-11) total = dx 2 . A comparison between E6-10 and E6-11 leads to du total = + vu = lam + turb . (E6-12) dr where =

Therefore, in turbulent flow, the turbulent shear stress consists of two components: du lam = = , as a result of fluid viscous effect; dr turb = ( u ' v ' ) , termed as Reynolds stress, as result of eddy transport.

Figure 6-11 Distribution of shear stress in turbulent pipe flow For turbulent pipe flow, the distribution of the two components of turbulent shear stress is shown in Figure 6-11. Governed by E6-11, the total shear stress is linear between the centreline and the wall. However, the contribution of viscous shear stress and Reynolds stress are significantly different from the pipe wall to the pipe centreline. As shown in Figure 6-11, near the pipe wall, viscous shear stress dominates while at the pipe centreline, Reynolds stress dominates and viscous shear stress is negligible. Therefore, the structure of turbulent pipe flow may be divided into three parts, as shown in Figure 6-12. 1. Viscous sublayer, where lam >> turb . In the viscous sublayer, the viscous effects are dominant, and the flow is laminar. The mean velocity, u , is zero at the wall and increases rapidly with r. 2. Buffer layer or Overlap layer, where lam turb Since both turbulent and laminar stresses are acting, this layer demonstrates properties of both laminar and viscous flow. 3. Turbulent core or Outer layer, where lam << turb Viscous stresses are negligible, resulting in almost constant velocity and little shear. This zone occupies 80-90% of the cross-sectional area of the pipe.

Prepared by Dr Hongwei Wu, since 2006

Chapter 6 Page 10 of 19

Figure 6-12 Structure of turbulent pipe flow 6.5.3 Velocity profile of turbulent pipe flow For laminar flow, the velocity profile can be analytically derived as E6-6. However, it is impossible to do so for a turbulent flow so that we can only obtain semi-empirical results. As the importance of viscous shear stress and Reynolds stress varies in different layers (see Figures 6-11 and 6-12), it is expected that the velocity profile varies in various layers.

Viscous sublayer We define a friction velocity, u * = w . Please note that u * is not a real physical quantity but u * has the same unit of velocity. In the viscous sublayer, experimental data show that u y = u* (E6-13) * u where = / and y = R r . Equation E6-13 is only valid within the viscous sublayer, which is taken as yu * 5. (E6-14)

Therefore, Equation E6-14 is often used to estimate the thickness of viscous sublayer.

In the buffer layer and turbulent core Experimental data in these two layers can be described by the following empirical equation. u y = 2.5 ln u * + 5.4 (E6-15) * u yu * For the buffer layer: 5< 50 (E6-16)

For the turbulent core:

yu *

> 50

(E6-17)

The turbulent flow profile is shown in Figure 6-13. Experimental data are plotted along with the predictions using E6-13 and E6-15.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 11 of 19

u u*

yu *

Figure 6-13 Turbulent flow profile, modified from [1]

6.5.4 Velocity profile of turbulent pipe flow the power law For turbulent pipe flow, the viscous sublayer is generally thin and the flow structure is dominated by the turbulent core. In 1932, Nikuradse proposed the following empirical power law velocity profile:
r n (E6-18) = 1 , u max R where n is a function of Reynolds number, and umax is the velocity on the pipe centreline. For many practical fluid, n = 7 is a reasonable approximation. Certainly, E6-18 CANNOT be used near the wall (i.e., viscous sublayer). u
1

Figure 6-14 Correlation between n and Reynolds number [1]


Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 12 of 19

r R

u Vc

Figure 6-15 Correlation between n and Reynolds number, modified from [1] Figure 6-14 shows that the value of n increases with Reynolds number, indicating the stronger the turbulence, the higher the n value. Figure 6-15 shows the velocity profiles at various n values. It can be seen that at strong turbulence (with a high n value), the pipe flow has nearly uniform velocity profile, for example the profile at n = 10.
6.6 Friction factor 6.6.1 Definition Lets introduce a dimensionless number, friction factor f. For a pipe with a length of L and a diameter of D, friction factor (Darcy friction factor) is defined as P( D / L) f = (E6-19) V 2 / 2 To understand the meaning of the friction factor in E6-19, we recall that E6-5 applies for all flows therefore the wall stress can be written as R P 1 w = = P( D / L) (E6-20) 2 L 4 Substituting into E6-19, we have the definition of friction factor f as 4 w f = (E6-21) V 2 / 2 Fundamentally, the friction factor f is the ratio of wall shear stress and inertial force of the flow. E6-19 and E6-21 is generic for laminar or turbulent flows. With the introduction of friction factor f, rewriting E6-19, the pressure drop can be calculated as L V 2 P = f (E6-22) D 2

or in head form
hL ,major = f L V2 D 2g

(E6-22a)
Chapter 6 Page 13 of 19

Prepared by Dr Hongwei Wu, since 2006

E6-22 (or E6-22a) is very important for the engineering design and operation of pipe systems. It shows that the pressure drop of a fully-developed pipe flow is a function of three parameters, i.e., the friction factor f, pipe geometry (L/D) and dynamic pressure V 2 / 2 (or velocity head V 2 / 2 g ). Therefore, pipe flow problems are converted to find f for the flows under different conditions. 6.6.2 Friction factor for laminar pipe flow For laminar pipe flow, according to E6-7, we know P = 32 VL / D 2 This is for laminar flow only! If we divided both sides by the dynamic pressure (
2

V 2
2

) and

L , we have D

64 P( D / L) 32VL / D D = = 64 2 2 VD = Re . V / 2 V / 2 L Therefore, for laminar flow, the friction factor is proportional to f = 64 . Re 1 , i.e. Re (E6-23)

Darcy friction factor and Fanning friction factor To make it hard for engineers, historically two friction factors are in common use. One is called Darcy friction factor as shown in E6-19 and E6-21, often appears in mechanical and civil engineering books. The other is called Fanning friction factor, which often appears in chemical engineering books. The Fanning friction factor is defined as P( D / L) f fanning = (E6-19a) 4( V 2 / 2)

i.e.

V 2 / 2 f Darcy= 4 f fanning .

f fanning =

(E6-21a)

If Fanning friction factor is used, the following equation should be used to calculate the pressure drop of pipe flows L V 2 P = 4 f Fanning (E6-22a) D 2 For laminar pipe flow 16 f Fanning = . (E6-23a) Re When use engineering books, one must be very careful about which friction factor is used. In this unit, we use the value of Darcy friction factor defined in (E6-19) and (E6-21).
6.6.3 Friction factor for turbulent flow Equations E6-19, E6-21 and E6-22 are applicable to turbulent pipe flow. However, for turbulent flow, it is impossible to analytically derive the friction factor f, which can ONLY be obtained from experimental data. In addition, most pipes, except glass tubing, have rough surfaces. The pipe surface roughness is quantified by a dimensionless number, relative pipe roughness ( / D ), where is pipe roughness and D is pipe diameter. For laminar pipe flow, the flow is dominated by viscous effects hence surface roughness is not a consideration.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 14 of 19

However, for turbulent flow, the surface roughness may protrude beyond the laminar sublayer and affect the flow to a certain degree. Therefore, the friction factor f can be generally written as a function of Reynolds number Re and pipe relative roughness
f = (Re,

: (E6-24)

For smooth pipes For turbulent flow in smooth pipes with Re < 105, Blasius [4] calculated the friction factor for a large variety of pipe-flow experiments and found that the friction factor f is indeed only a function of Reynolds number, 0.316 f = 1/ 4 (E6-25) Re

For rough pipes Depending on the depth of surface protrusions, one would expect that it is possible to determine the following three regimes, 1. If the surface protrusions are within the viscous sublayer, the pipe can be treated as hydraulically smooth so that the flow is dependent only on the Reynolds number, i.e. E6-24 can be simplified to f = (Re) , i.e., E6-25. 2. If the surface protrusions extend into just the buffer layer, the flow is dependent on both the Reynolds number and the relative roughness. Details of E6-24 needs to be obtained from experimental investigations. 3. For larger protrusions into the turbulent core, the effect of surface roughness is so strong that the flow is dependent only on pipe roughness, i.e. f = ( ) . D
6.6.4 Moody chart As discussed in the previous two subsections, the friction factor for turbulent flow is a function of Reynolds number and pipe roughness, described in the generic equation E6-24. In order to find the detailed equation of E6-24, large quantities of experiments have been carried out using pipes of various relative surface roughness under different flow conditions.

Figure 6-16 shows the functional dependence of f on Re and D and is called the Moody chart [5]. The Moody Chart covers all three flow regions: laminar, transitional and completely turbulent. Table 6-1 lists the typical roughness from various new, clean pipe surfaces. To use the Moody chart, we need to 1. Calculate the relative roughness ( D ) and the Reynolds number ( Re = VD ) for the flow under consideration. 2. Follow the line for the particular roughness value until we locate the Reynolds number, and read the friction factor from the left axis. 3. Take extra care when reading the Reynolds number axis, which goes up to 108. In particular, note the pattern of values along the axis, e.g., 105 2 4 6 8 106, etc. If a value 8.5 falls into the region, the exponent needs to be read TO THE LEFT (i.e. 105), not to the right. This is a significant source of error when determining friction factors, and hence leading to possible faulty pipe designs.

Prepared by Dr Hongwei Wu, since 2006

Chapter 6 Page 15 of 19

Table 6-1 Equivalent roughness for new pipes [5] Equivalent roughness, Pipe Feet Millimetres Glass, plastic Smooth Smooth Drawn tubing (e.g. copper) 0.000005 0.0015 Commercial steel or wrought iron 0.00015 0.045 Galvanised iron 0.0005 0.15 Cast iron 0.00085 0.26 Concrete 0.001 0.01 0.3 3 Riveted steel 0.003 0.03 0.9 9.0 It should be noted that the Moody chart is to help us understand the fundamentals and practice hand solutions. Working with the Moody chart directly can sometimes be tedious and errorprone. Most modern engineers prefer to use computer programs to solve flow problems. In this case, besides the analytical solutions from laminar flow, empirical correlations are developed to simulate partially regions of the Moody chart with good accuracy, including 64 (E6-26a) For laminar flow: f = Re 0.316 For turbulent flow in smooth pipes, Re < 105 f = 1/ 4 (E6-26b) Re /D 1 2.51 = 2.0 log + (E6-26c) For the entire turbulent region: 3.7 Re f f
6.7 Revisit of Figure 6-3 Region I: Laminar pipe flow The test section length is x. For laminar pipe flow, according to E6-22, we have the pressure gradient as 2 P 1 V (E6-27) = f x D 2 The fluid velocity V can be calculated from flow rate Q and pipe cross-section area A (which is determined by pipe diameter D) , Q 4Q (E6-28) V= = A D 2 According to E6-23, for laminar pipe flow (with the substitution of V using E6-28), 64 64 16D = = f = Re VD Q (E6-29)

Substituting E6-28 and E6-29 into E6-27, yielding 2 16D 1 1 4Q 128 P 1 V ( 2 )2 = = f = Q Q D 2 D x D 2 D 2 128 is a constant (assumed as c1), therefore, we have In experiments leading to Figure 6-3, D 2 P (E6-30) = c1Q x In Region I, E6-30 therefore holds. The pressured gradient should be proportional to the flow rate. This was indeed validated by the experimental data in Figure 6-3.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 16 of 19

f =

P ( D / L ) 1 V 2 2

Figure 6-16 The Moody Chart, adapted from [1] which the original data were from [5]
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 17 of 19

Region III: Turbulent flow E6-27 and E6-28 are still applicable. We need to use the friction factor in turbulent flow. 1. For rough pipes From the Moody chart, one can see that when Reynolds number is high enough, the flow is wholly turbulent flow. The friction factor f is only a function of relative pipe roughness, independent of Reynolds number. The same pipe was used in experiment therefore the relative pipe roughness should be a constant. Consequently, the friction factor should be constant (denoted as c),
f = ( ) = c D Substituting E2-28 and E2-31 into E6-27, we have 2 P 4Q 8 f 1 1 1 V = f ( 2 )2 = 2 5 Q 2 = f x D D 2 D D 2 .

(E6-31)

In the experiments leading to Figure 6-3,

P = c2Q 2 x Therefore the pressure gradient is proportional to the square of the flow rate.

8 f is a constant (denoted as c2), we have 2 D5

(E6-30)

2. For smooth pipes Blasius equation E6-25 (also E6-26b) can be used to calculate the friction factor f for VD turbulent flow in smooth pipes. Substituting Re = and E6-28 into E6-25, we have

1/ 4 D 0.316 0.316( ) 0.25 f = 1/ 4 = = 0.316 1/ 4 1/ 4 4 Q Re (D ) (V ) 1/ 4

(E6-32)

Substituting E2-28 and E2-32 into E6-27, we have


2 D P 1 V = f = 0.316 4 x D 2 1/ 4

1 4 ( 2 ) 2 Q1.75 . 2 D D
1/ 4

D 4 2 1 In the experiments leading to Figure 6-3, 0.316 4 2 D ( D 2 ) is a constant (denoted as c3), we have P (E6-33) = c3Q1.75 x Hence the pressure gradient is proportional to the flow rate to ~1.8 power.

6.8 Noncircular Conduits

Many practical engineering situations use conduits of non-circular cross section. Certainly, the details of the flows in noncircular conduits depend on the exact cross-sectional shape. However, the discussion on flow in round pipe can be applied in noncircular conduits with slight modifications.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 18 of 19

We introduce an easy-to-use parameter, called hydraulic diameter, DH, defined as 4A (E6-34) DH = P where A is the cross-sectional area of the noncircular pipe and P is the wetted perimeter. Therefore, for noncircular pipe, we used DH in P( DH / L) friction factor , f = V 2 / 2 VDH Re = , Reynolds number

relative roughness

DH

For round pipe of diameter D, we can calculate DH = D 2 D = D . Some typical value of DH is listed in Table 6-2. Table 6-2 Hydraulic diameter (DH) of pipe of typical cross-sectional shape Cross-sectional shape of pipe DH Square

Rectangular 2ab a+b

Concentric Annulus

D2 D1

References [1]. Munson BR, Young DF and Okiishi TH, Fundamentals of Fluid Mechanics, 4th Edition, John Wiley & Sons, Brisbane, 2002. [2]. Noel de Nevers, Fluid Mechanics for Chemical Engineers, 3rd Edition, McGraw-Hills Chemical Engineering Series, Sydney 2005. [3]. Reynolds O. An experimental investigation of the circumstances which determine whether the motion of water shall be direct or sinuous and of the law of resistance in parallel channels, Philosophy Transactions of the Royal Society, 174, 1883. [4]. White FM, Viscous Fluid Flow, McGraw-Hill, New York, 1979. [5]. Moody LF, Friction factors for pipe flow, Transactions of the ASME, 1944; 66: 671.
Prepared by Dr Hongwei Wu, since 2006 Chapter 6 Page 19 of 19

You might also like