You are on page 1of 14

Single Phase Flow Calculations

for
UBD Applications

Prepared By

Garry A. Gregory, P.Eng.

© May 2002

Notes for a Professional Development Course


presented in

Calgary, Alberta, Canada

December 4 - 6, 2002

NEOTECHNOLOGY CONSULTANTS LTD.


510, 1701 Centre Street N.W.
Calgary, Alberta, Canada T2E 7Y2
Tel: (403) 277-6688 Fax: (403) 277-6687
Internet: www.neotec.com
Disclaimer

Although all of the information contained in these Short Course Notes is believed to be accurate at the time it
is prepared, it is presented without representation or warranty of any kind. Neither Neotechnology Consultants
Ltd. nor the author shall assume any liability of any kind whatsoever, either collectively or individually, arising
from the use or application of any of the technology, descriptions, or expressed opinions contained herein.
Table of Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2. The Mechanical Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

3. Friction Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

4. The Total Energy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

5. The Stepwise Calculation Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

6. Simultaneous Calculation of Pressure and Temperature Profiles . . . . . . . . . 10

7. Literature References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
SINGLE PHASE FLOW FOR APPLICATIONS

1. Introduction

The mathematical derivations of the fundamental relationships governing the steady state single phase
flow of compressible and incompressible fluids in pipes are well documented in most introductory texts on
fluid mechanics (e.g. Streeter, 1962; Massey, 1979) and will not be repeated here. The purpose of these Notes
is rather to demonstrate how these relationships are applied to the design and analysis of single phase gas and
liquid production and transmission systems, with particular emphasis on hydrocarbon fluids.

2. The Mechanical Energy Equation

For steady state flow of a fluid in a pipe with a uniform cross section, the mechanical energy balance,
in differential form, is given by,

(2-1)

or, in integrated form,

(2-2)

where )P = overall pressure loss (psi or Pa)


D = fluid density (lb/ft3 or kg/m3)
g = gravitational acceleration (32.2 ft/s2 or 9.81 m/s2)
gc = dimensional conversion constant (32.2 ft-lbm/lbf - s2 or 1.0)
)z = vertical elevation change (ft or m)
V = average fluid velocity (ft/s or m/s)
" = velocity profile factor (0.5 for laminar flow, typically 0.8 to 0.9 for turbulent flow)
) Pf = irreversible (i.e. friction) pressure loss (psi or Pa)

In performing the integration to get from Equation (2-1) to Equation (2-2), it must be assumed that
the fluid is incompressible. For most liquids, the effect of pressure on the density is negligible for most
practical purposes, and the assumption is easily satisfied. However, even for the flow of gases, if the analysis
is based on the Stepwise Calculation procedure that was described earlier in Chapter 3 of these Notes, it is
easily demonstrated that the assumption can be justified if the step size is selected carefully. This point will
be discussed in greater detail a little later in this Chapter.

The first term on the right hand side of Equation (2-2) is the contribution to the overall pressure
change that results from a change in the potential energy of the fluid (i.e. changes that result from the
elevation of the fluid). As such, it is a reversible change, and can thus account for either a pressure loss
(uphill flow) or a pressure gain (downhill flow).

The second term on the right hand side of Equation (2-2) is the contribution to the overall pressure
Single Phase Flow for UBD Applications (Revised 05-02) Page 2

change that results from changes in the kinetic energy of the fluid. This is often referred to as the acceleration
term, since it is related to velocity changes. This too represents a reversible change, and it can result from
either changes in the density of the fluid or in the diameter of the pipe.

The final term on the right hand side of Equation (2-2) is a pressure loss that accounts for all of the
energy that is degraded to internal energy as a result of the flow. It is an irreversible term, and is most
commonly known as the friction loss.

We could thus re-write Equation (2-2) in simpler terms as,

(2-3)

where )P = overall pressure loss in the system (psi or kPa)


)PPE = pressure loss due to potential energy changes (i.e. elevation effects) (psi or kPa)
)PKE = pressure loss due to kinetic energy changes (i.e. acceleration effects) (psi or kPa)
) Pf = pressure loss due to irreversible effects (i.e. friction losses) (psi or kPa)

This is a particularly useful form of the mechanical energy balance equation, since it clearly illustrates the
partitioning of the overall pressure loss into its contributing components.

For example, )PPE can also be written as,

(2-4)

where )L = the length of the pipe section (ft or m)


2 = the angle of the pipe from horizontal in the direction of flow (degrees)

For a purely horizontal pipe, 2 = 0 o ; there is no elevation change, and, it is clear from Equation
(2-4) that )PPE = 0 in such cases.

For incompressible flow (e.g. most liquids) in a constant diameter pipe, )PKE = 0. In fact, even for
compressible fluids (e.g. gases), it is only in relatively rare cases that kinetic energy effects are significant.
These invariably involve high rate flow at relatively low pressure, such that the pressure loss is a significant
fraction of the overall pressure. In most systems, however, one is normally justified in ignoring this term
altogether.

The final term, )Pf , must almost always be considered, since it represents an irreversible loss. At
low to moderate flow rates, it can be small, possibly even negligible in systems with significant elevation
changes. We will examine this term in greater detail very shortly.

It is important to note at this point that Equations (2-1) through (2-4) are perfectly general, and they
can be applied to any section of constant diameter pipe. In other words, these relationships apply equally to
both pipelines and wells. It is only the relative magnitudes of the terms that are affected by the orientation
of the pipe.

3. Friction Losses
Single Phase Flow for UBD Applications (Revised 05-02) Page 3

Except for purely laminar flow, no theoretical way exists to evaluate )Pf . Calculation of this term
for turbulent flow is generally based on the Fanning (1893) friction factor, f, which is defined by the relation,

(3-1)

where D = pipe inside diameter (ft or m)


f = Fanning (1893) friction factor (dimensionless)

Equation (3-1) can be rearranged to give,

(3-2)

which allows us to compute )Pf directly, provided that we know f .

Unfortunately, Equation (3-1) simply defines f , and does not give us any indication of how to
evaluate it when )Pf is not known. Note, however, that as a defining equation, it is an inherently exact
relationship. This is an important point that is often forgotten, but it has significant implications, especially
for gas flow in pipelines and wells, as we shall see shortly.

Fortunately, dimensional analysis can be used to show that f can be adequately described as a function
of a dimensionless group, known familiarly as the Reynolds number, or Re, defined as,

(3-3)

where: : = fluid viscosity (lb/ft-s or Pa.s)

Experimental work by Nikuradse (1933) confirmed that there was also a systematic dependence of
f on the relative pipe roughness, k/D. There is often some confusion regarding exactly what the absolute pipe
roughness, k actually represents. It is not the actual dimension of pitting or other surface roughness
characteristics (although it is related to those), but it also includes the effects of other factors such as weld
beads. It is, in fact, based on experiments performed by Nikuradse (1933), who coated the inside of pipes
with carefully graded (i.e. approximately uniform diameter) sand particles. Pressure drop measurements
subsequently made in those sand coated pipes were compared with the results of similar experiments
performed in uncoated pipes. The effective roughness for the uncoated pipes could then be expressed in terms
of an equivalent sand grain roughness, or, the diameter of the particles in a graded sand coating that would
give the same pressure loss under the given flow conditions.

The resulting correlation between those parameters is usually presented in the form of the familiar
Moody (1944) diagram, shown in Figure 1. This is a demonstrably accurate correlation, and it applies to the
flow of any incompressible Newtonian fluid in a circular cross section pipe.
Single Phase Flow for UBD Applications (Revised 05-02) Page 4

Figure 1

Moody Diagram - Friction Factor vs Reynolds Number and Relative Roughness

In the case of laminar flow (i.e. Re < 2100), it is a relatively simple matter to derive the theoretical result
from first principles that,

(3-4)

In laminar flow, therefore, the friction loss depends only on the Reynolds number and it is independent of the
pipe roughness.

We will digress briefly here to note that there is another friction factor that is known generally as the
Darcy-Weisbach (1845, 1854) friction factor. It is also frequently used in the literature (especially by
mechanical engineers), and is defined by,

(3-5)

Comparison of Equation (3-5) with Equation (3-1) leads to the simple result that,
Single Phase Flow for UBD Applications (Revised 05-02) Page 5

(3-6)

and thus, for laminar flow,

(3-7)

For consistency, we use only the Fanning (1893) friction factor throughout these Notes. You should
be aware, however, the Moody diagrams appear in textbooks and in the published literature in terms of both
f and f D . Significant errors are thus possible if you mismatch the numerical value of the friction factor
with its defining equation to compute a friction pressure loss. The distinction between the two is always
clearly made in the laminar flow region of the diagram.

For computational purposes, it is useful to have an analytical expression for f when the flow is
turbulent, rather than having to use the Moody diagram. One of the simplest of these is the Drew et al (1932)
equation,
(3-8)

Note that since this equation does not include the absolute pipe roughness k, it is strictly applicable only for
smooth pipes, and it is, in fact, a good representation of the line corresponding to k = 0 on the Moody
diagram.

A more complex expression that does take into account the roughness of the pipe is the Colebrook-
White (1939) equation,

(3-9)

One often hears it said that the Colebrook-White equation should be the preferred expression for computing
the Fanning friction factor, because of its excellent overall agreement with the standard Moody diagram. This
reaction is not too surprising, since the Colebrook-White equation has, in fact, been used to generate the
Moody diagram that appears in almost every fluid mechanics text book! On the other hand, it is also an
implicit relationship in f , and it is thus not a particularly convenient equation to use routinely, since an
iterative solution for f is required. This is not a major problem, as the equation is not particularly complex
and the solution converges relatively quickly using only standard numerical procedures.

In many typical pipeline calculations, however, f must be evaluated hundreds or even thousands of
times. Gregory and Fogarasi (1985) noted that using an iterative solution to evaluate f can add between 10%
and 15% to the overall time required to perform a particular pipeline calculation, compared to using an
explicit relationship, where iterations are not required. Those authors evaluated numerous explicit equations
for f which had appeared in the published literature and concluded that many of them were reasonable
alternatives to the Colebrook-White equation. For example, over a range of Reynolds numbers and relative
pipe roughness values that cover most of the expected operating range for typical pipelines and wells, average
agreement between values of f computed using an explicit equation were well within 2 - 3% of those
computed using the Colebrook-White equation. By comparison, if the pipe relative roughness was varied
only ± 10%, the variation in f computed by the latter was typically about ± 5%. Considering that one would
seldom ever know the true relative pipe roughness to within ± 10%, many of the explicit equations were thus
Single Phase Flow for UBD Applications (Revised 05-02) Page 6

considered to give acceptable accuracy.

In particular, however, Gregory and Fogarasi (1985) recommended the use of the Chen (1979)
equation because of its consistently excellent agreement with the Colebrook-White equation over the full
ranges of Reynolds numbers and relative pipe roughness that were investigated. This relationship can be
written as follows,

(3-10)

(3-11)

(3-12)

While this may appear to be a rather formidable set of equations, compared to Equation (3-9), in a number
of benchmark computer runs it was actually about 12% faster, on average, and gave values that were
generally indistinguishable from those generated by the Colebrook-White equation.

For clean, new, carbon steel pipe, the absolute roughness k is generally taken to be about 0.0018
inches (0.046 mm). For some liquid service, conservative practice suggests that a larger value (say, 0.0025
inches, or 0.065 mm) should be used to account for the long term effect of scaling and corrosion over the life
of the pipeline. However, one should also expect to modify the design value of k if the pipeline is to be
pigged regularly (i.e. “pig burnished” pipe is typically less rough than new pipe) , or if the pipe is coated
internally. For example, a paper by Worthingham et al (1994) reported results from a study involving 48-inch
diameter internally coated pipe where the effective roughness was observed to be as low as 0.00025 inches
(0.0064 mm), or about 14% of the normal value. The relative economics of coating the pipe versus additional
compression costs become the deciding factor in any given case.

For performance analysis studies, the value of k that should be used will depend to some degree on
how long the line has been in service, and whether it has been used for corrosive and/or waxy fluids. Typical
values of k for a number of materials are shown in Table 1.

Before leaving this Section, it is worth noting that Equation (3-2) can be written in a number of
different ways that may be simpler to use, depending on the form of the data and information that you have
to work with.

For example, the mass flow rate, W is defined by,

(3-13)

Table 1
Single Phase Flow for UBD Applications (Revised 05-02) Page 7

Typical Values of Absolute Roughness for Various Pipe Materials (Moody, 1944)

Absolute Absolute
Pipe Material Roughness Roughness
(inches) (mm)

Drawn tubing .00006 .0015


Commercial Steel .0018 .046
Fiberglass .00021 .0053
Wrought Iron .0018 .046
Asphalted cast iron .0048 .122
Galvanized Iron .006 .152
Cast iron .0102 .259
Concrete .012 - .12 .305 - 3.05
Wood stave .0102 - .036 .259 - 2.59
Riveted steel .036 - .36 .914 - 9.14

where W = mass flow rate (lb/s or kg/s)

If Equation (3-13) is used to replace the velocity V in Equation (3-2), we obtain,

(3-14)

Similarly, the mass flux, or mass velocity, G is defined by the relation,

(3-15)

where G = mass flux (lb/ft2-s or kg/m2-s)

If Equation (3-15) is substituted into Equation (3-2), we obtain,

(3-16)

Finally, the volumetric flow rate, Q is given by the expression,

(3-17)
Single Phase Flow for UBD Applications (Revised 05-02) Page 8

where Q = volumetric flow rate (ft3/s or m3/s)

Substitution of Equation (3-17) into Equation (3-2) yields the result,

(3-18)

Any of Equations (3-14), (3-16), or (3-18) are thus equivalent to Equation (3-2).

4. The Total Energy Equation

The total energy equation is an expression of the balance that must exist at steady state between heat
that is transferred to or from the flowing fluid to the pipe surroundings and other energy changes that occur
within the system.

Assuming that there are no unusual energy factors (e.g. chemical reactions, electromagnetic fields),
for some particular length of pipe, the total energy equation may be written as,

(4-1)

where q = rate of heat transfer from the fluid to the pipe surroundings (BTU/s or kJ/s)
W = mass flow rate of the fluid (lb/s or kg/s)
H = specific enthalpy of the fluid (BTU/lb or kJ/kg)
V = average velocity if the fluid (ft/s or m/s)
z = elevation of the length of pipe (ft or m)
J = energy conversion factor (778 ft-lbf/BTU or 1000 J/kJ)
" = velocity profile factor (0.5 for laminar flow, typically 0.8 to 0.9 for turbulent flow)
g = gravitational acceleration (32.2 ft/s2 or 9.81 m/s2)
gc = dimensional conversion constant (32.2 ft-lbm/lbf-s2 or 1.0)

The subscripts 1 and 2 refer to conditions at the upstream and downstream ends of the pipe section
respectively. The mass flow rate of the fluid can be computed from the equation,

(4-2)

where D = inside diameter of the pipe (ft or m)


D = average density of the fluid in the pipe section (lb/ft3 or kg/m3)

The rate of heat transfer between the flowing fluid and the pipe surroundings can be computed on
the basis of an overall heat transfer coefficient, using the familiar relation,

(4-3)
Single Phase Flow for UBD Applications (Revised 05-02) Page 9

where Do = pipe outside diameter (ft or m)


)L = the length of the pipe section (ft or m)
Uo = overall heat transfer coefficient (BTU/ft2-s or kJ/m2-s)
Tf = average temperature of the flowing fluid (°F or °C)
Ts = average temperature of the pipe surroundings (°F or °C)

It is clear from Equation (4-3) that the magnitude of the heat transfer term is directly proportional to
the magnitude of the overall heat transfer coefficient. The latter is a measure of the ease with which thermal
energy is transferred between the flowing fluid and its surroundings.

The estimation of Uo is usually based on a resistance-in-series concept. Every source of resistance


to heat transfer between the flowing fluid and the surroundings (e.g. laminar film adjacent to the inside wall
of the pipe, the pipe wall, insulation, outside film, transition zone(s) in the surroundings) contributes in some
degree to the value of Uo that should be used. An insulated pipe will have a higher resistance to heat transfer
than an uninsulated one, and the heat transfer coefficient will be lower as a result.

Calculation methods for Uo are described in general in most basic text books on heat transfer. Some
fairly comprehensive and detailed discussions on calculating this parameter for pipelines and wells have been
presented elsewhere by Gregory (1991b, 1991c)

The first term on the right hand side of Equation (4-1) is the change in enthalpy of the fluid as is
passes from point 1 to point 2 in the pipe section of interest. The original term in the overall energy balance
is in fact the Internal Energy of the fluid. It is a simple matter to show, however, that when no work is being
done to or by the fluid, this reduces to a simple enthalpy change.

The second term on the right hand side of Equation (4-1) is the change in Kinetic Energy. For
constant diameter pipe. this term is always negligible for liquids, and unless the pressure change is a
significant fraction of the absolute pressure, it is also insignificant for gases. The exceptions typically occur
with gases at relatively low pressure, or in a blowout situation. In such cases, this term can dominate and,
in fact, be the controlling factor in the overall pressure loss.

The last term on the right hand side of Equation (4-1) is the Potential Energy change. It is a
reversible term and ranges in importance from negligible in near horizontal systems, or gas lines at low
pressure, to 95% or more of the total pressure loss in oil wells or other near vertical systems.

5. The Stepwise Calculation Procedure

The most accurate way to perform calculations for flow in pipes is to use the Stepwise calculation
procedure. This is a differential type of calculation in which the pipe system being considered is treated as
a connected sequence of pipe segments. There can be as many calculation segments as required or practical,
but it is assumed that each segment satisfies the following criteria:

(i) the pipe has a constant angle or slope


(ii) the pipe has a constant diameter and relative roughness
(iii) the total mass flow within the pipe is constant
(iv) the nature of the pipe surroundings is unchanged
(v) the mechanical properties of the pipe (diameter, wall thickness, material) are unchanged
(vi) the thermal properties of the pipe (conductivity, heat capacity) are unchanged

Changes in any of the above are thus assumed to occur at the junction between two segments.

While requirements (ii) through (vi) above can usually be met very closely, if not exactly,
Single Phase Flow for UBD Applications (Revised 05-02) Page 10

requirement (i) is generally only satisfied approximately unless a very large number of pipe sections is
considered. In most cases however, this has a minimal effect on the results of any flow calculations, provided
that only small variations in slope are permitted within a given section. Many multiphase models, for
example, have dependencies which vary only with the magnitude of the rise and the drop, rather than with
the actual slope of the pipe. However, most of the sophisticated mechanistic models for multiphase flow do
take into account the actual slope; when such models are to be used, the drilling profile should be specified
as accurately as possible.

For a given pipe calculation segment, it is further assumed that the pressure and temperature are
known at one end. One then proceeds as follows:

(1) Estimate the overall pressure drop across the segment, and thus the average (i.e. midpoint)
pressure. It is assumed that the average temperature can similarly be estimated with reasonable
accuracy.

(2) Use appropriate P-V-T behaviour and transport property model(s) to determine all require fluid
properties at the assumed average conditions in the segment.

(3) Use an appropriate fluid flow model to compute the overall pressure drop, based on the P-V-T
behaviour and fluid transport properties that were estimated in (2).

(4) Compare the pressure drop computed in (3) with that assumed in (1). If the two agree to within
an acceptable tolerance, the procedure can be terminated and be assumed to have converged..
Otherwise, use the calculated pressure drop as a new estimate, and repeat (1) to (4) as required.

The downstream pressure and temperature that are calculated for the given segment then represent
the upstream conditions for the next calculation segment. The iterative procedure is then repeated for that
segment, and the calculations continue in this way until the entire piping system has been traversed. In
practice, the number of steps or calculation segments required can be quite large, especially for deep wells
or wells with relatively high pressure gradients. Clearly, this is a procedure that had to wait until we routinely
had access to high speed computing facilities!

In the Stepwise calculation procedure, one always try to control the step size such that the pressure
changes are small enough to ensure that the mid-point fluid properties will be representative of those
everywhere in the calculation segment (i.e. the pressure gradient is essentially linear). This ensures that the
even gases can be treated as if they were incompressible fluids, and exactly the same procedure can be used
as for liquids.

6. Simultaneous Calculation of Pressure and Temperature Profiles

The Stepwise Calculation procedure described above forms the basis of a rigorous and theoretically
sound algorithm for calculating, simultaneously, the pressure and temperature profiles in the flowing system.
Since there are two unknowns (pressure and temperature), two independent equations are required. Two such
relations are Equation (2-2), the mechanical energy equation, and Equation (4-1), the total energy equation
.

Within each calculation segment, one first assumes the values of the average pressure and
temperature. All relevant fluid properties are evaluated at these assumed conditions, using appropriate
methods, and are substituted into Equations (2-2) and (4-1) to determine whether or not those relations are
satisfied. An iterative procedure is followed, using successive trial values of pressure and temperature until
agreement is obtained for both equations to satisfactory convergence criteria. The calculations then move
along to the next pipe segment until the entire system has been traversed.
Single Phase Flow for UBD Applications (Revised 05-02) Page 11

In specific cases, where either the temperature profile is assumed to be known or the temperature can
be assumed to be constant, the pressure profile can be computed solely from the mechanical energy equation.
This, of course, reduces both the amount of data required and the time required to perform the calculations.

Once again, the point must be made that both the equations and the calculation procedure are valid
for any pipe orientation. The procedure is thus completely general and can be used for all sections of a well
(vertical, deviated, and horizontal). Clearly, however, one must have a detailed knowledge of the drilling
profile for the flow system to properly account for the potential energy term in the calculated profiles.

7. Literature References

Chen, N. H., "An Explicit Equation for Friction Factor in Book Company, Inc., New York
Pipe", Ind. Eng. Chem. Fund., vol. 18, p. 296 (1979)
Weisbach, J., “Lehrbuch der Ingenieur - und Maschinen-
Colebrook, C. F., "Turbulent Flow in Pipes with mechanik”, Braunschweig:, F. Vieweg und Sohn (1845)
Particular Reference to the Transition Region Between the
Smooth and Rough Pipe Laws", J. Inst. Civ. Eng., vol. 11, Worthington, R. G., Asante, B., Carmichael, G. A., and
p. 133 (1939) Dunsmore, T., “Cost Study Justifies Internal Coating on
48-inch Gas Lines”, Oil & Gas J., p. 80, May 30 (1994)
Colebrook, C. F., and White, C. M., “Experiments with
Fluid Friction in Roughened Pipes”, Proc. Royal Soc.
(London), 161-A, p. 367 (1937)

Darcy. H., “About Experimental Research Concerning the


Movement of Water Through Pipes”, Compte Rend., vol
38, No. 11, p. 1109 (1854)

Drew, T. B., Koo, E. C., and McAdams, W. H., Trans.


A.I.Ch.E., vol. 28, p. 56 (1932)

Fanning, J. T., A Practical Treatise on Hydraulic and


Water Supply Engineering, D. Van Nostrand, New York,
(1893)
Calgary, AB Canada (1989)

Gregory, G. A., "Estimation of the Overall Heat Transfer


Coefficient for the Calculation of Pipeline Heat
Loss/Gain", Technical Note No. 3, Neotechnology
Consultants Ltd., Calgary, AB Canada (1991 b)

Gregory, G. A., "Estimation of the Overall Heat Transfer


Coefficient for Calculating Heat Loss/Gain in Flowing
Wells", Technical Note No. 4, Neotechnology Consultants
Ltd., Calgary, AB Canada (1991 c)

Gregory, G. A., and Fogarasi, M., "Alternate to Standard


Friction Factor", Oil & Gas Jour., pg. 120, April 1 (1985)

Massey, B. S., Mechanics of Fluids, 4th Edition, Van


Nostrand Reinhold Company, New York, NY (1979)

Moody, L. F., “Friction Factors for Pipe Flow”, Trans.


A.S.M.E., vol. 66, p. 671 (1944)

Nikuradse, J., “Flow Law in Rough Pipes (in German),


VDI-Forschungsheft, p. 361 (1933)

Streeter, V. L., Fluid Mechanics, 3rd ed., McGraw-Hill

You might also like