You are on page 1of 17

WELLBORE PERFORMANCE

For designing oil well equipment and optimizing well production conditions is vitally important to
production engineers to understand wellbore flow performance. Wellbore performance analysis
creates a relationship among tubular size, wellhead and bottom-hole pressure, fluid properties, and
fluid production rate.
The objective of wellbore flow performance is generally to forecast the pressure as a function of
position between the bottomhole location and surface.
Wellbore flow has several broad categories, depending on the :

 Flow geometry
 Fluid properties
 Flow rate
Initially, the flow in a wellbore is either single phase or multiphase; in most production wells, the flow
is multiphase, with at least two phases (e.g., gas and liquid) present, but also some production wells
and most injection wells experience single-phase flow,

 FLOW GEOMETRY :-
The flow geometry in the wellbores is usually flow through a circular pipe. Although flow in an
annular space, such as between tubing and casing, sometimes occurs.

 FLUID PROPERTIES :-
In describing wellbore flow performance, the fluid properties must consider both their PVT behavior
and rheological characteristics.

 FLOW RATE :-
Flow in a wellbore may be either laminar or turbulent, and this will strongly influence the flow
behavior.

SINGLE PHASE LIQUID FLOW :-


When the wellhead pressure is above the bubble-point pressure of the oil, single-phase liquid flow
exists in an oil well, which is not reality. To understand the concept of fluid flow in oil wells where
multiphase flow usually dominates, it is suitable to start from single phase liquid.
Consider a fluid flowing from point 1 to point 2 in a tubing string of length L and height Δz.
The first law of thermodynamics yields the following equation for pressure drop:

where
ΔP=pressure drop, lbf/ft2
P1=pressure at point 1, lbf/ft2
P2=pressure at point 2, lbf/ft2
g=gravitational acceleration, 32.17 ft/s2
gc=unit conversion factor, 32.17 lbm-ft/lbf-

Figure 1 Flow along a tubing string

s2
ρ=fluid density, lbm/ft3
Δz=elevation increase, ft
u=fluid velocity, ft/s
fF=Fanning friction factor
L=tubing length, ft
D=tubing inner diameter, ft

Flow along a tubing string (Courtesy The Production Technology)

The first term in the right-hand side of the equation represents the change in pressure due to change in
elevation.
The second term in the right-hand side of the equation represents the change in pressure due to change
in kinetic energy.
The third term in the right-hand side of the equation represents the change in pressure due to change
in friction.
The Fanning friction factor (fF) can be determined by Reynolds number and relative roughness. The
Reynolds number is the ratio of the inertial forces to the viscous forces in a flowing fluid, and for flow
in a circular pipe is given by

or in U.S. field units as


where
NRe = Reynolds number
q= fluid flow rate, bbl/day
ρ= fluid density, lbm/ft3
d= tubing inner diameter, in.
μ= fluid viscosity, cp

Single-phase flow can be described as being either laminar or turbulent, depending on the Reynolds
number, NRe. The kind of flow, will strongly influence the velocity profile in the pipe, the frictional
pressure drop behavior, and the dispersion of solutes contained in the fluid, among other factors; all of
these elements must be considered at times in production operations.

In laminar flow, fluid moves in a smooth distinct layer, with no or little mixing with the bulk flow
direction. However turbulent flow is described by eddy current which causes fluctuation of velocity
components in all directions.
The conversion from laminar to turbulent flow in circular pipes generally occurs at a Reynolds
number of 2100, depending on the pipe roughness, entrance conditions, and other factors, this value
can vary. For laminar flow NRe<2000, and the fF is inversely proportional to the Reynolds number, or

For turbulent flow where NRe>2100, the fF can be estimated using empirical correlations.

SINGLE PHASE GAS FLOW :-


The first law of thermodynamics regulates gas flow in tubing. because the variation in tubing diameter
is insignificant in most gas wells, the effect of kinetic energy change is negligible. With no shaft work
device installed along the tubing string, the first law of thermodynamics produces the following
mechanical balance equation:

Because dZ=cos θdL, , and


Eq. (4.7) can be rewritten as
This differential equation governs the gas flow in tubing. This equation is difficult to solve
analytically because the temperature T can be approximately expressed as a linear function of
length L through geothermal gradient, the compressibility factor z is a function of pressure P and
temperature T. Luckily, the pressure P at length L is not a strong function of temperature and
compressibility factor.
Approximate solutions to Eq. (4.8) have been required and used in the natural gas industry.

AVERAGE TEMPERATURE AND COMPRESSIBILITY FACTOR


METHOD :-
If single average values of temperature and compressibility factor over the entire tubing length can be
assumed, Eq. becomes

By separation of variables, Eq. can be integrated over the full length of tubing to yield

where

Eqs take the following forms when U.S. field units (qsc in Mscf/d), are used (Katz et al., 1959):

and

The Darcy–Wiesbach (Moody) friction factor fM can be found in the conventional manner for a given
tubing diameter, wall roughness, and Reynolds number. However, if one assumes fully turbulent flow,
which is the case for most gas wells, then a simple empirical relation may be used for typical tubing
strings (Katz and Lee, 1990):

Guo and Ghalambor (2002) used the following Nikuradse friction factor correlation for fully turbulent
flow in rough pipes:
Because the average compressibility factor is a function of pressure itself, a numerical technique such
as Newton–Raphson iteration is required to solve above Eq. for bottom-hole pressure. This
computation can be performed automatically with the spreadsheet program Average TZ.xls. Users
need to input parameter values in the Input data section and run Macro Solution to get results

CULLENDER AND SMITH METHOD :-


Above Eq. can be solved for bottom-hole pressure using a fast numerical algorithm originally
developed by Cullender and Smith (Katz et al., 1959). Eq. (4.8) can be rearranged as

that takes an integration form of

In U.S. field units (qmsc in MMscf/d), Eq. has the following form:

If the integrant is denoted with symbol I, that is,

Eq. becomes
In the form of numerical integration, Eq. can be expressed as

where pmf is the pressure at the mid-depth. The Ihf, Imf, and Iwf are integrant is evaluated at phf, pmf,
and pwf, respectively. Assuming the first and second terms in the right-hand side of Eq. each
represents half of the integration, that is,

the following expressions are obtained:

Because Imf is a function of pressure pmf itself, a numerical technique such as Newton–Raphson
iteration is required to solve Eq. for pmf. Once pmf is computed, pwf can be solved numerically from Eq.
These computations can be performed automatically with the spreadsheet program Cullender-
Smith.xls. Users need to input parameter values in the Input Data section and run Macro Solution to
get results.

MULTIPHASE FLOW IN WELLS :-


Multiphase flow—the simultaneous flow of two or more phases of fluid—will occur in almost all oil
production wells, in many gas production wells, and in some types of injection wells. The TPR
equation for single-phase flow is not valid for multiphase oil wells. To analyze TPR of multiphase oil
wells rigorously, a multiphase flow model is required.
Multiphase flow is much more complicated than single-phase flow because of the variation of flow
regime (or flow pattern). Fluid distribution changes greatly in different flow regimes, which
significantly affects pressure gradient in the tubing.

FLOW REGIMES IN GAS-LIQUID FLOW :-


It is a description of the geometrical distribution of multiphase fluid moving through a pipe. Many
different terms are used to describe this distribution, the distinction between each one being
qualitative and somewhat arbitrary
A flow regime is determined by the velocity of the gas and the liquid phases and the relative amount
of gas and liquid at any given point in the flowstream. Multiphase flow in a vertical conduit is usually
represented by four basic flow regimes :
 Bubble Flow
 Slug Flow
 Churn Flow
 Annular Flow
 Wispy annular flow
 Mist flow

These flow regimes occur as a progression with increasing gas flow rate for a given liquid flow rate.
A brief description of these flow regimes is as follows :

1. BUBBLE FLOW :-
The tubing is almost completely filled with liquid. Free gas is present as small bubbles, rising in the
liquid. Liquid contacts the wall surface, and the bubbles serve only to reduce the density.
Numerous bubbles are observable as the gas is dispersed in the form of discrete bubbles in the
continuous liquid phase. The bubbles may vary widely in size and shape but they are typically nearly
spherical and are much smaller than the diameter of the tube itself

1. SLUG FLOW :-
At higher gas rates, the bubbles coalesce into larger bubbles, called Taylor bubbles, that eventually fill
the entire pipe cross section. Between the large gas bubbles are slugs of liquid that contain smaller
bubbles of gas entrained in the liquid.
With increasing gas void fraction, the proximity of the bubbles is very close such that bubble collide
and coalesce to form larger bubbles, which are similar in dimension to the tube diameter. These
bubbles have a characteristic shape similar to a bullet with a hemisphere nose with a blunt tail end.
They are commonly referred to as Taylor bubbles. Taylor bubbles are separated from one another by
slugs of liquid, which may include small bubbles. Taylor bubbles are surrounded by a thin liquid film
between them and the tube wall, which may flow downward due to the force of gravity, even though
the net flow of fluid is upward.

2. CHURN FLOW :-
With a further increase in gas rate, the larger gas bubbles become unstable and collapse, resulting in
churn flow, a highly turbulent flow pattern with both phases dispersed. Churn flow is characterized by
oscillatory, up-and-down motion of the liquid.
Increasing the velocity of the flow, the structure of the flow becomes unstable with the fluid travelling
up and down in an oscillatory fashion but with a net upward flow. The instability is the result of the
relative parity of the gravity and shear forces acting in opposing direction on the thin film of liquid of
Taylor bubbles. This flow pattern is in fact an intermediate region between the slug flow and annular
flow regimes.
3. ANNULAR FLOW :-
At very high gas rates, gas becomes the continuous phase, with liquid flowing in an annulus coating
the surface of the pipe and with liquid droplets entrained in the gas phase.
Once the interfacial shear of the high velocity gas on the liquid film becomes dominant over gravity,
the liquid is expelled from the center of the tube and flows as a thin film on the wall (forming an
annular ring of liquid) while the gas flows as a continuous phase up the center of the tube. The
interface is distributed by high frequency waves and ripples. In addition, liquid may be entrained in
the gas core as small droplets, so much so that the fraction of the liquid entrained may become similar
to that in the film. This flow regime is particularly stable and is the desired flow pattern for the two-
phase pipe flows.

4. WISPY-ANNULUAR FLOW
When the flow rate is further increased, the entrained droplets may become transient coherent
structure as clouds or wisps of liquid in the central vapor core.

5. MIST FLOW
At very high gas flow rates, the annular film is thinned by the shear of the gas core on the interface
until it becomes unstable and is destroyed, such that all the liquid in entrained as droplets in the
continuous gas phase, analogous to the inverse of the bubbly flow regime.

Figure 2 Two-phase flow pattern in vertical upward flow

Two-phase flow pattern in vertical upward flow (Courtesy The Production Technology)
TUBING PERFORMANCE RELATIONSHIP CURVE :-
A well’s production is dependent on the mechanical configuration of the wellbore, the fluid
properties, the reservoir conditions and several other factors. There are several ways to determine the
well’s performance.
One way is the tubing performance relation (TPR) curve, also known as the tubing performance curve
(TPC), the vertical lift performance and the outflow performance curve.

Figure 3 Two simplified depiction of gas well reservoirs


system

Two simplified depictions of gas well-reservoirs system ( Courtesy The Production Engineering )

There are two versions of the TPR curve used in practice. The first depicts the relation between the
pressure drop of the well and the flow rate at the well head shown in figure 1.1, i.e. the top site flow
rate. This is the version used in this report. The second depicts the relation between the bottom hole
pressure and the top site flow rate. This second curve just adds the well head pressure to the pressure
drop to give the actual pressure at the bottom of the well. It is merely the first curve plus a constant.

In general the pressure drop is determined using the mechanical energy equation for flow between
two points of the system,

𝑃1 𝑢12 𝑃2 +𝑢22
+ 𝑔𝑧1 + = + 𝑔𝑧2 + 𝑊 + 𝐸1
𝜌 2 𝜌 2

With the variables p, u, z, ρ, g ,W and El representing respectively pressure, flow rate, depth, density,
gravitational acceleration, work on the fluid and the irreversible energy loses. Where the subscripts 1
and 2 denote the location within the system. For the setup in this report the mass conservation and
momentum conservation are used instead of the mechanical energy equation. This may be done since
the mechanical energy equation is equivalent to the steady state momentum conservation equation C,
The fluid inside the gas well often contains liquid, i.e. there is two-phase flow. This results in a specific
shape for the TPR curve. The pressure drop over the well needs to be large enough to compensate the
forces working on the fluid. At low flow rates the gravitational force is large due to the large ratio of
liquid. At high flow rates the frictional force between the fluid and the well is large. This results in what
is often called a J-curve as in figure below.

Figure 4- Standard shape of the TPR curve ( Courtesy The Production Engineering )

TPR Models
Numerous TPR models have been developed for analyzing multiphase flow in vertical pipes. Brown
(1977) presents a thorough review of these models. TPR models for multiphase flow wells fall into
two categories: (1) homogeneousflow models and (2) separated-flow models. Homogeneous models
treat multiphase as a homogeneous mixture and do not consider the effects of liquid holdup (no-slip
assumption). Therefore, these models are less accurate and are usually calibrated with local operating
conditions in field applications. The major advantage of these models comes from their mechanistic
nature. They can handle gas-oilwater three-phase and gas-oil-water-sand four-phase systems. It is
easy to code these mechanistic models in computer programs.

Separated-flow models are more realistic than the homogeneous-flow models. They are usually given
in the form of empirical correlations. The effects of liquid holdup (slip) and flow regime are considered.
The major disadvantage of the separated flow models is that it is difficult to code them in computer
programs because most correlations are presented in graphic form.

HOMOGENOUS-FLOW MODELS
Numerous homogeneous-flow models have been developed for analyzing the TPR of multiphase wells
since the pioneering works of Poettmann and Carpenter (1952). Poettmann–Carpenter’s model uses
empirical two-phase friction factor for friction pressure loss calculations without considering the effect
of liquid viscosity. The effect of liquid viscosity was considered by later researchers including Cicchitti
(1960) and Dukler et al. (1964). A comprehensive review of these models was given by Hasan and
Kabir (2002). Guo and Ghalambor (2005) presented work addressing gas-oil-water-sand four-phase
flow. Assuming no slip of liquid phase, Poettmann and Carpenter (1952) presented a simplified gas-
oil-water threephase flow model to compute pressure losses in wellbores
by estimating mixture density and friction factor. According to Poettmann and Carpenter, the following
equation can be used to calculate pressure traverse in a vertical tubing when the acceleration term is
neglected:

𝑘̅ 𝛥ℎ
𝛥𝑃 = ( 𝜌̅ + )
𝜌̅ 144

where
𝛥𝑃 = pressure increment, psi
𝜌̅ = average mixture density (specific weight), lb/ft3
𝛥ℎ = depth increment, ft
And
𝑓2𝐹 𝑞𝑜2 𝑀2
𝑘̅ = 5
7.4137∗1010 𝐷

where
f2F = Fanning friction factor for two-phase flow
𝑞𝑜 = oil production rate, stb/day
M = total mass associated with 1 stb of oil
D = tubing inner diameter, ft

The average mixture density 𝜌̅ can be calculated by

𝜌̅1 + 𝜌̅2
𝜌̅ =
2

where
𝜌̅1 = mixture density at top of tubing segment, lb/ft3
𝜌̅2 = mixture density at bottom of segment, lb/ft3

The mixture density at a given point can be calculated based on mass flow rate and volume flow rate:

𝑀
𝜌=
𝑉𝑚
Where

𝑀 = 350.17 ( ɣ𝑜 + 𝑊𝑂𝑅 ɣ𝑊 ) 𝐺𝑂𝑅 𝜌𝑎𝑖𝑟 ɣ𝑔

14.7 𝑇 𝑍
𝑉𝑚 = 5.615 ( 𝐵𝑜 + 𝑊𝑂𝑅 𝐵𝑤 ) + (𝐺𝑂𝑅 − 𝑅𝑆 ) ( )( )( )
𝑃 520 1.0

and where

ɣ𝑜 = oil specific gravity, 1 for freshwater


WOR = producing water–oil ratio, bbl/stb
ɣ𝑊 = water-specific gravity, 1 for freshwater
GOR = producing gas–oil ratio, scf/stb
𝜌𝑎𝑖𝑟 = density of air, lbm=ft3
ɣ𝑔 = gas-specific gravity, 1 for air
Vm = volume of mixture associated with 1 stb of oil, ft3
Bo = formation volume factor of oil, rb/stb
Bw = formation volume factor of water, rb/bbl
Rs = solution gas–oil ratio, scf/stb
p = in situ pressure, psia
T = in situ temperature, 8R
z = gas compressibility factor at p and T.

If data from direct measurements are not available, solution gas–oil ratio and formation volume factor
of oil can be estimated using the following correlations:

𝑃 100.0125𝐴𝑃𝐼
𝑅𝑠 = 𝛾𝑔[ ]
18 100.00091𝑡

ɣ𝑔 0.5
𝐵𝑂 = 0.9759 + 0.00012 [ 𝑅𝑠 ( ) + 1.25 𝑡 ]1.2
ɣ𝑜

where t is in situ temperature in 8F. The two-phase friction factor f2F can be estimated from a chart
recommended by Poettmann and Carpenter (1952). For easy coding in computer programs, Guo and
Ghalambor (2002) developed the following correlation to represent the chart:

𝑓2𝐹 = 101.44 −2.5 log(𝐷𝜌𝑣 )


where (Drv) is the numerator of Reynolds number representing inertial force and can be formulated as
−5
1.4375 ∗ 10 𝑀𝑞𝑜
(𝐷𝜌𝑣 ) =
𝐷

Because the Poettmann–Carpenter model takes a finite difference form, this model is accurate
for only short depth incremental Δh. For deep wells, this model should be used in a piecewise
manner to get accurate results (i.e., the tubing string should be ‘‘broken’’ into small segments
and the model is applied to each segment). Because iterations are required to solve Eq. for
pressure, a computer spreadsheet program Poettmann-CarpenterBHP.xls has been developed.

SEPARATE-FLOW MODELS
A number of separated-flow models are available for TPR calculations. Among many others are the
Lockhart and Martinell icorrelation (1949), the Duns and Ros correlation (1963), and the Hagedorn
and Brown method (1965). Based on comprehensive comparisons of these models, Ansari et al.
(1994) and Hasan and Kabir (2002) recommended the Hagedorn–Brown method with modifications
for near-vertical flow. The modified Hagedorn–Brown (mH-B) method is an empirical correlation
developed on the basis of the original work of Hagedorn and Brown (1965). The modifications
include using the no-slip liquid holdup when the original correlation predicts a liquid holdup value
less than the noslip holdup and using the Griffith correlation (Griffith and Wallis, 1961) for the bubble
flow regime. The original Hagedorn–Brown correlation takes the following form:

𝑑𝑃 𝑔 ̅ 𝑢𝑚 2
2𝑓𝐹 𝜌 𝛥 (𝑢𝑚 2 )
𝑑𝑧
= 𝑔𝑐
𝜌̅ 𝑔𝑐 𝐷
+ 𝜌̅ 2 𝑔𝑐 𝛥𝑧

which can be expressed in U.S. field units as

𝑑𝑃 𝑓𝐹 𝑀𝑡 2 𝛥 (𝑢𝑚 2 )
144 = 𝜌̅ 10
+ 𝜌̅
𝑑𝑧 7.4137 ∗ 10 𝐷5 𝜌̅ 2 𝑔𝑐 𝛥𝑧

where
Mt = total mass flow rate, lbm/d
𝜌̅ = in situ average density, lbm/ft3
um = mixture velocity, ft/s
and

𝜌̅ = 𝑦𝐿 𝜌𝐿 + (1 − 𝑦𝐿 )𝜌𝐺

𝑢𝑚 = 𝑢𝑆𝐿 + 𝑢𝑆𝐺

where
𝜌𝐿 = liquid density, lbm/ft3
𝜌𝐺 =in situ gas density, lbm/ft3
𝑢𝑆𝐿 uSL = superficial velocity of liquid phase, ft/s
𝑢𝑆𝐺 =superficial velocity of gas phase, ft/s
The superficial velocity of a given phase is defined as the volumetric flow rate of the phase divided by
the pipe cross-sectional area for flow. The third term in the right-hand side of Eq. represents pressure
change due to kinetic energy change, which is in most instances negligible for oil wells.

Obviously, determination of the value of liquid holdup yL is essential for pressure calculations. The
mH-B correlation uses liquid holdup from three charts using the following dimensionless numbers:

Liquid velocity number, NvL :

4 𝜌𝐿
𝑁𝑣𝐿 = 1.938 𝑢𝑆𝐿 √
𝜎

Gas velocity number, NvG :

4 𝜌𝐺
𝑁𝑣𝐺 = 1.938 𝑢𝑆𝐿 √
𝜎

Pipe diameter number, ND:

𝜌𝐿
𝑁𝐷 = 120.827𝐷 √
𝜎

Liquid viscosity number, NL:

4 1
𝑁𝐿 = 0.15726 µ𝐿 √ 𝜎3
𝜌𝐿

where
D = conduit inner diameter, ft
𝜎 = liquid–gas interfacial tension, dyne/cm
µ𝐿 = liquid viscosity, cp
µ𝐺 = gas viscosity, cp

The first chart is used for determining parameter (CNL) based on NL. We have found that this chart
can be replaced by the following correlation with acceptable accuracy:

(𝐶𝑁𝐿 ) = 10ɣ
Where

Y = −2.69851 + 0.15841 𝑋1 – 0.55100 𝑋1 2 + 0.54785 𝑋1 3 − 0.12195 𝑋1 4

And

𝑋1 = log [(𝑁𝐿 ) + 3]

Once the value of parameter (CNL) is determined, it is used


for calculating the value of the group
𝑁𝑣𝐿 𝑝0.1 (𝐶𝑁𝐿 )
𝑁𝑣𝐺 0.575 𝑝𝑎 0.1 𝑁𝐷

where
p is the absolute pressure at the location where pressure gradient is to be calculated, and
pa is atmospheric pressure.
The value of this group is then used as an entry in the second chart to determine parameter
(yL=c). We have found that the second chart can be represented by the
following correlation with good accuracy:
ɣ𝐿
= −0.10307 + 0.61777[ log (X2) + 6] − 0.63295[ log (X2) + 6]2 + 0.29598[ log (X2)6]3
𝛹
− 0: 0401[ log (X2) + 6]4

Where
𝑁𝑣𝐿 𝑝0.1 (𝐶𝑁𝐿 )
𝑋2 =
𝑁𝑣𝐺 0.575 𝑝𝑎 0.1 𝑁𝐷

According to Hagedorn and Brown (1965), the value of parameter c can be determined from the third
chart using a value of group

𝑁𝑣𝐺 𝑁𝐿 0.38
𝑁𝐷 2.14
𝑁𝑣𝐺 𝑁𝐿 0.38
We have found that for 𝑁𝐷 2.14
> 0.01 the third chart can be replaced by the following correlation
with acceptable
accuracy:

Ψ = 0:91163 – 4.82176X3 + 1,232.25 X32 - 22,253.6 X33 + 116174,3 X43

Where

𝑁𝑣𝐺 𝑁𝐿 0.38
𝑋3 =
𝑁𝐷 2.14

𝑁𝑣𝐺 𝑁𝐿 0.38
However, Ψ=1.0 should be used for 𝑁𝐷 2.14
< 0.01. Finally, the liquid holdup can be calculated by

𝑦𝐿
𝑦𝐿 = 𝛹( )
𝛹

The Reynolds number for multiphase flow can be calculated by :

2.2 ∗ 10−2 𝑚𝑡
𝑁𝑅𝑒 =
𝐷 𝜇𝐿 𝑦𝐿 µ𝐺 (1−𝑦𝐿 )

where mt is mass flow rate. The modified mH-B method uses the Griffith correlation for the bubble-
flow regime. The bubble-flow regime has been observed to exist when

𝜆 𝐺 < 𝐿𝐵

Where
𝑢𝑠𝐺
𝜆𝐺 =
𝑢𝑚
And

𝑢𝑚 2
𝐿𝐵 = 1.071 − 0.2218 ( )
𝐷

which is valid for 𝐿𝐵 > 0.13. When the 𝐿𝐵 value given by Eq. (4.45) is less than 0.13, 𝐿𝐵 = 0.13
should be used. Neglecting the kinetic energy pressure drop term, the Griffith correlation in U.S. field
units can be expressed as

𝑑𝑃 𝑓𝐹 𝑚𝐿 2
144 = 𝜌̅ 10
𝑑𝑧 7.4137 ∗ 10 𝐷5 𝜌𝐿 𝑦𝐿 2

where 𝑚𝐿 is mass flow rate of liquid only. The liquid holdup in Griffith correlation is given by the
following expression:

1 𝑢𝑚 𝑢𝑚 2 𝑢𝑠𝐺
𝑦𝐿 = 1 − 2 ( 1 + 𝑢𝑠
√(1 + 𝑢𝑠
) −4 𝑢𝑠
)

where 𝑢𝑠 =0:8 ft=s. The Reynolds number used to obtain the friction factor is based on the in situ
average liquid velocity, that is,

2.2 ∗ 10−2 𝑚𝐿
𝑁𝑅𝑒 =
𝐷 𝜇𝐿

You might also like