100% found this document useful (1 vote)
283 views

Pressure Drop Calculation

1) The document discusses pressure drop calculations in cylindrical pipes. It presents the Darcy-Weisbach equation which relates pressure drop to fluid properties, pipe diameter, flow rate, and friction factor. 2) Flow can be laminar, critical, or turbulent depending on the Reynolds number. Equations are provided relating friction factor to Reynolds number for laminar and turbulent flow in smooth pipes. 3) For rough pipes, a roughness Reynolds number is defined and charts called Moody diagrams relate friction factor to both Reynolds number and relative roughness. Empirical equations are discussed that model the transition between smooth and rough pipe flow.

Uploaded by

David Lambert
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
283 views

Pressure Drop Calculation

1) The document discusses pressure drop calculations in cylindrical pipes. It presents the Darcy-Weisbach equation which relates pressure drop to fluid properties, pipe diameter, flow rate, and friction factor. 2) Flow can be laminar, critical, or turbulent depending on the Reynolds number. Equations are provided relating friction factor to Reynolds number for laminar and turbulent flow in smooth pipes. 3) For rough pipes, a roughness Reynolds number is defined and charts called Moody diagrams relate friction factor to both Reynolds number and relative roughness. Empirical equations are discussed that model the transition between smooth and rough pipe flow.

Uploaded by

David Lambert
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 8

Pressure Drop Calculation

In a cylindrical pipe of uniform diameter D, flowing full, the pressure loss due to viscous effects
Δp is proportional to length L and can be characterized by the Darcy–Weisbach equation:[2]

where the pressure loss per unit length Δp/L (SI units: Pa/m) is a function of:

ρ, the density of the fluid (kg/m3);


D, the hydraulic diameter of the pipe (for a pipe of circular section, this equals the
internal diameter of the pipe; otherwise D ≈ 2√A/π for a pipe of cross-sectional area A)
(m);
⟨v⟩, the mean flow velocity, experimentally measured as the volumetric flow rate Q per
unit cross-sectional wetted area (m/s);
fD, the Darcy friction factor (also called flow coefficient λ[3][4]).

For laminar flow, the friction factor is inversely proportional to the Reynolds number alone
(fD = 64/Re ) which itself can be expressed in terms of easily measured or published physical
quantities (see section below). Making this substitution the Darcy-Weisbach equation is rewritten
as

where

μ is the dynamic viscosity of the fluid (Pa·s = N·s/m2 = kg/(m·s));


Q is the volumetric flow rate, used here to measure flow instead of mean velocity
according to Q = π/4 D2⟨v⟩ (m3/s).

Note that this laminar form of Darcy–Weisbach is equivalent to the Hagen–Poiseuille equation,
which is analytically derived from the Navier–Stokes equations.

Head-loss form
The head loss Δh (or hf) expresses the pressure loss due to friction in terms of the equivalent
height of a column of the working fluid, so the pressure drop is

where

Δh is the head loss due to pipe friction over the given length of pipe (SI units: m);[b]
g is the local acceleration due to gravity (m/s2).

It is useful to present head loss per length of pipe (dimensionless):

where L is the pipe length (m).

Therefore, the Darcy–Weisbach equation can also be written in terms of head loss:[5]

In terms of volumetric flow

The relationship between mean flow velocity ⟨v⟩ and volumetric flow rate Q is

where:

Q is the volumetric flow (m3/s),


A is the cross-sectional wetted area (m2).

In a full-flowing pipe of diameter D,

Then the Darcy–Weisbach equation in terms of Q is

Shear-stress form
The mean wall shear stress τ in a pipe or open channel is expressed in terms of the Darcy–
Weisbach friction factor as[6]

The wall shear stress has the SI unit of pascals (Pa).

Darcy friction factor


Figure 1. The Darcy friction factor versus Reynolds number for 10 < Re < 108 for smooth pipe
and a range of values of relative roughness ε/D . Data are from Nikuradse (1932, 1933),
Colebrook (1939), and McKeon (2004).

The friction factor fD is not a constant: it depends on such things as the characteristics of the pipe
(diameter D and roughness height ε), the characteristics of the fluid (its kinematic viscosity ν
[nu]), and the velocity of the fluid flow ⟨v⟩. It has been measured to high accuracy within certain
flow regimes and may be evaluated by the use of various empirical relations, or it may be read
from published charts. These charts are often referred to as Moody diagrams, after L. F. Moody,
and hence the factor itself is sometimes erroneously called the Moody friction factor. It is also
sometimes called the Blasius friction factor, after the approximate formula he proposed.

Figure 1 shows the value of fD as measured by experimenters for many different fluids, over a
wide range of Reynolds numbers, and for pipes of various roughness heights. There are three
broad regimes of fluid flow encountered in these data: laminar, critical, and turbulent.

Laminar regime

For laminar (smooth) flows, it is a consequence of Poiseuille's law (which stems from an exact
classical solution for the fluid flow) that

where Re is the Reynolds number

and where μ is the viscosity of the fluid and

is known as the kinematic viscosity. In this expression for Reynolds number, the characteristic
length D is taken to be the hydraulic diameter of the pipe, which, for a cylindrical pipe flowing
full, equals the inside diameter. In Figures 1 and 2 of friction factor versus Reynolds number, the
regime Re < 2000 demonstrates laminar flow; the friction factor is well represented by the above
equation.[c]
In effect, the friction loss in the laminar regime is more accurately characterized as being
proportional to flow velocity, rather than proportional to the square of that velocity: one could
regard the Darcy–Weisbach equation as not truly applicable in the laminar flow regime.

In laminar flow, friction loss arises from the transfer of momentum from the fluid in the center of
the flow to the pipe wall via the viscosity of the fluid; no vortices are present in the flow. Note
that the friction loss is insensitive to the pipe roughness height ε: the flow velocity in the
neighborhood of the pipe wall is zero.

Critical regime

For Reynolds numbers in the range 2000 < Re < 4000, the flow is unsteady (varies grossly with
time) and varies from one section of the pipe to another (is not "fully developed"). The flow
involves the incipient formation of vortices; it is not well understood.

Turbulent regime

Figure 2. The Darcy friction factor versus Reynolds number for 1000 < Re < 108 for smooth
pipe and a range of values of relative roughness ε/D . Data are from Nikuradse (1932, 1933),
Colebrook (1939), and McKeon (2004).

For Reynolds number greater than 4000, the flow is turbulent; the resistance to flow follows the
Darcy–Weisbach equation: it is proportional to the square of the mean flow velocity. Over a
domain of many orders of magnitude of Re (4000 < Re < 108), the friction factor varies less than
one order of magnitude (0.006 < fD < 0.06). Within the turbulent flow regime, the nature of the
flow can be further divided into a regime where the pipe wall is effectively smooth, and one
where its roughness height is salient.

Smooth-pipe regime

When the pipe surface is smooth (the "smooth pipe" curve in Figure 2), the friction factor's
variation with Re can be modeled by the Kármán–Prandtl resistance equation for turbulent flow
in smooth pipes[3] with the parameters suitably adjusted

The factors 1.930 and 1.90 are phenomenological; these specific values provide a fairly good fit
to the data.[7] The product Re√fD (called the "friction Reynolds number") can be considered, like
the Reynolds number, to be a (dimensionless) parameter of the flow: at fixed values of Re√fD ,
the friction factor is also fixed.

In the Kármán–Prandtl resistance equation, fD can be expressed in closed form as an analytic


function of Re through the use of the Lambert W function:

In this flow regime, many small vortices are responsible for the transfer of momentum between
the bulk of the fluid to the pipe wall. As the friction Reynolds number Re√fD increases, the
profile of the fluid velocity approaches the wall asymptotically, thereby transferring more
momentum to the pipe wall, as modeled in Blasius boundary layer theory.

Rough-pipe regime

When the pipe surface's roughness height ε is significant (typically at high Reynolds number),
the friction factor departs from the smooth pipe curve, ultimately approaching an asymptotic
value ("rough pipe" regime). In this regime, the resistance to flow varies according to the square
of the mean flow velocity and is insensitive to Reynolds number. Here, it is useful to employ yet
another dimensionless parameter of the flow, the roughness Reynolds number[8]

where the roughness height ε is scaled to the pipe diameter D.

Figure 3. Roughness function B vs. friction Reynolds number R∗. The data fall on a single
trajectory when plotted in this way. The regime R∗ < 1 is effectively that of smooth pipe flow.
For large R∗, the roughness function B approaches a constant value. Phenomenological functions
attempting to fit these data, including the Afzal[9] and Colebrook–White[10] are shown.

It is illustrative to plot the roughness function B:[11]

Figure 3 shows B versus R∗ for the rough pipe data of Nikuradse,[8] Shockling,[12] and
Langelandsvik.[13]
In this view, the data at different roughness ratio ε/D fall together when plotted against R∗,
demonstrating scaling in the variable R∗. The following features are present:

 When ε = 0, then R∗ is identically zero: flow is always in the smooth pipe regime. The
data for these points lie to the left extreme of the abscissa and are not within the frame of
the graph.
 When R∗ < 5, the data lie on the line B(R∗) = R∗; flow is in the smooth pipe regime.
 When R∗ > 100, the data asymptotically approach a horizontal line; they are independent
of Re, fD, and ε/D .
 The intermediate range of 5 < R∗ < 100 constitutes a transition from one behavior to the
other. The data depart from the line B(R∗) = R∗ very slowly, reach a maximum near R∗ =
10, then fall to a constant value.

A fit to these data in the transition from smooth pipe flow to rough pipe flow employs an
exponential expression in R∗ that ensures proper behavior for 1 < R∗ < 50 (the transition from the
smooth pipe regime to the rough pipe regime):[9][14][15]

This function shares the same values for its term in common with the Kármán–Prandtl resistance
equation, plus one parameter 0.34 to fit the asymptotic behavior for R∗ → ∞ along with one
further parameter, 11, to govern the transition from smooth to rough flow. It is exhibited in
Figure 3.

The Colebrook–White relation[10] fits the friction factor with a function of the form

[d]

This relation has the correct behavior at extreme values of R∗, as shown by the labeled curve in
Figure 3: when R∗ is small, it is consistent with smooth pipe flow, when large, it is consistent
with rough pipe flow. However its performance in the transitional domain overestimates the
friction factor by a substantial margin.[12] Colebrook acknowledges the discrepancy with
Nikuradze's data but argues that his relation is consistent with the measurements on commercial
pipes. Indeed, such pipes are very different from those carefully prepared by Nikuradse: their
surfaces are characterized by many different roughness heights and random spatial distribution of
roughness points, while those of Nikuradse have surfaces with uniform roughness height, with
the points extremely closely packed.

Calculating the friction factor from its parametrization

See also: Darcy friction factor formulae

For turbulent flow, methods for finding the friction factor fD include using a diagram, such as the
Moody chart, or solving equations such as the Colebrook–White equation (upon which the
Moody chart is based), or the Swamee–Jain equation. While the Colebrook–White relation is, in
the general case, an iterative method, the Swamee–Jain equation allows fD to be found directly
for full flow in a circular pipe.[5]

Direct calculation when friction loss S is known

In typical engineering applications, there will be a set of given or known quantities. The
acceleration of gravity g and the kinematic viscosity of the fluid ν are known, as are the diameter
of the pipe D and its roughness height ε. If as well the head loss per unit length S is a known
quantity, then the friction factor fD can be calculated directly from the chosen fitting function.
Solving the Darcy–Weisbach equation for √fD ,

we can now express Re√fD :

Expressing the roughness Reynolds number R∗,

we have the two parameters needed to substitute into the Colebrook–White relation, or any other
function, for the friction factor fD, the flow velocity ⟨v⟩, and the volumetric flow rate Q.

Confusion with the Fanning friction factor

The Darcy–Weisbach friction factor fD is 4 times larger than the Fanning friction factor f, so
attention must be paid to note which one of these is meant in any "friction factor" chart or
equation being used. Of the two, the Darcy–Weisbach factor fD is more commonly used by civil
and mechanical engineers, and the Fanning factor f by chemical engineers, but care should be
taken to identify the correct factor regardless of the source of the chart or formula.

Note that

Most charts or tables indicate the type of friction factor, or at least provide the formula for the
friction factor with laminar flow. If the formula for laminar flow is f = 16/Re , it is the Fanning
factor f, and if the formula for laminar flow is fD = 64/Re , it is the Darcy–Weisbach factor fD.

Which friction factor is plotted in a Moody diagram may be determined by inspection if the
publisher did not include the formula described above:
1. Observe the value of the friction factor for laminar flow at a Reynolds number of 1000.
2. If the value of the friction factor is 0.064, then the Darcy friction factor is plotted in the
Moody diagram. Note that the nonzero digits in 0.064 are the numerator in the formula
for the laminar Darcy friction factor: fD = 64/Re .
3. If the value of the friction factor is 0.016, then the Fanning friction factor is plotted in the
Moody diagram. Note that the nonzero digits in 0.016 are the numerator in the formula
for the laminar Fanning friction factor: f = 16/Re .

The procedure above is similar for any available Reynolds number that is an integer power of
ten. It is not necessary to remember the value 1000 for this procedure—only that an integer
power of ten is of interest for this purpose.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy