Sei sulla pagina 1di 67

211

Chapter 8

OIL AND GAS TRANSPORT

SANJAY KUMAR and GEORGE V. CHILINGARIAN

MTRODUCTION

Inasmuch as it is usually much cheaper to use pipelines than to use barges,


tankers, trucks, etc., complex transcontinental piping systems have been developed.
Some of the major advantages of using pipelines include:
(1) Economy.
(2) Reliability under almost all conditions (e.g., adverse weather, breakdowns,
non-availability of tankers, etc.).
(3) Control: An installed pipeline can usually handle a wide range of flow rates.
(4) Continuity of flow, which is highly desirable in modern continuous-flow oil
refineries, because a minimum of storage facilities are required at either end.
The designs of simple systems are briefly discussed here. The fundamental concepts
involved are very useful and their knowledge is absolutely essential for any practic-
ing petroleum engineer.

FUNDAMENTALS OF FLOW IN PIPES

Reynolds applied dimensional analysis to the phenomenon of the transition from


laminar to turbulent flow. He concluded that it occurred at a fixed value of a
dimensionless group, which is called the Reynolds number ( N R e )in his honor:

NRe = dVp/p = d V / v = inertia forces/viscous forces (8-1)

where d = diameter of the conduit through which the fluid is flowing, V = velocity

V = O AT PIPE WALL
I

LAMINAR FLOW TURBULENT FLOW

Fig. 8-1.Velocity profiles in a circular conduit for laminar flow (left) and turbulent flow (right).
212

2
of the fluid, p = density of the fluid,’ p = dynamic viscosity of fluid, and
Y = kinematic viscosity of fluid.
For laminar or viscous flow, N,, < 2100; for transitional or intermediate flow,
2100 < N R e < 10,000 3 , and for turbulent flow, N,, > 10,000.3
For cases where the flow is not through a cylindrical pipe (the cross-section is
other than circular), the concept of effective diameter has been developed:

d, =4R, (8-2)
where hydraulic radius, R , = (area of flow)/(wetted perimeter).
If the cross-section of a pipe is square, for example, cross-sectional area of
flow = a’,

d a2 a
wetted perimeter = 4 a , and R , = = -= -. Thus: d , = 4 R , = a .
4 4a 4
In the case of circular cross section, area of flow, A = r d 2 / 4 , wetted perimeter
=rd, R , = ( m d 2 / 4 ) / ( r d )= d / 4 , and d , = 4 R , = d .
The schematic diagrams of velocity profiles for laminar and turbulent flows are
shown in Fig. 8-1.
Allowable working pressure of pipeline
The allowable working pressure of a pipeline, p a , is equal to:
St’Fe ( d o - d , )
Pa = (8-3)
do
where S = specified minimum yield strength, psi; t’ = temperature derating factor,
dimensionless; F = design factor (construction type); e = longitudinal joint factor
(i.e., anomaly due to weld seam); it is equal to 1.0, except for butt-welded
ASTM-A53, API-5L ( = 0.6), fusion-welded A 134 and A 139 ( = 0.8), and spiral-
welded A 211 (= 0.8); d , = inside diameter of pipe, in.; d o = outside diameter of
pipe, in.

The specific weight, y , is the weight per unit volume and can be expressed in terms of lb/cu ft. The
density, p’, which is mass per unit volume and is equal to y / g , where g is the gravitational acceleration
( = 32.174 ft-sec-’), can be expressed in terms of slugs/cu ft. For example, pure water, which has a
specific weight of 62.4 lb/cu ft, has a density, p ’ , of 1.94 ( = 62.4/32.17) slugs/cu ft. In other words, the
mass is attracted by the earth with a force of magnitude p’g. Inasmuch as 1 slug = 32.17 lb,, a density of
1 slug/ft3 = 32.17 lb,/ft3. The density expressed in lb,/ft3 will be designated by p throughout this
book.
Inasmuch as p = shear stress/shear strain = T/(dV/dy), the units of dynamic viscosity are Ibfsec/ft2
or slugs/ft-sec. 1 poise = 100 centipoises = 1dyne-sec/cm2 = 1 g/cm-sec = 0.00209 slug/ft-sec = 0.00209
lb,sec/ft’ = 0.0673 Ib,/ft-sec. 1 slug/ft-sec = 479 poises. 1 stoke -100 centistokes -1 cm2/sec =
0.001076 ft‘/sec.
Value of 3100-4000 is commonly assigned to this limit.
213

Horsepower

The hydraulic horsepower, P,, is equal to 0.000017 X bbl/D X psig = 0.00053 X


gal/min X psig.

Friction

Obviously, friction is associated with any kind of flow. Pipes are defined as
smooth, if the relative roughness k, (= r/di) is < ( c = absolute surface
roughness in inches and di = inside diameter of pipe in ft).
The Fanning equation for steady state flow in uniform circular pipes which are
full of liquid under isothermal conditions is as follows:

where A p , = pressure loss due to friction, lb/ft2; f = Fanning. friction factor,


dimensionless; I = length of pipe, ft; V = fluid velocity, ft/sec; y = fluid specific
weight, lb/ft3; p = density (mass/unit volume), lb,/ft3; g = gravitational accelera-
tion, 32.2 ft/sec2; g, = dimensional constant, 32.2 lb,-ft/lb-sec2; and d = diameter
of pipe, ft.
If A p , is expressed in psi, the eq. 8-4 becomes:

The Fanning friction factor, f , is equal to:

f = - for laminar flow, '


16
NRe

f = 0.1419( N R e ) -0.3192 for intermediate flow,

f = 4 log- NReJT for partially turbulent flow, and


1.4126

According to Hagen and Poiseuille, at N R e ranging from 2000 to 2300, f = 64/NR,. Thus, extreme
care must be exercised on using values of f available in the literature, because they may be multiples of
the f given in Fig. 8-2. The equation f = 64/( V d p / p ) is widely accepted.
214

1 = 4 log- NReJT
-
JT 1.255

for fully turbulent flow (Prandtl-Khrmhn formula)


For rough pipes, using the absolute roughness, E in microinches, the above relation-
ships are slightly modified:

1 2.51
(8-9)

for intermediate flow (Colebrook's relation),

1 3.71di
sr
- = 4 log-
NRe
for turbulent flow (Prandtl-Khrmhn equation), (8-10)

f=f, + 0.68NRef? - (Supino formula), (8-11)


i i i )

where f, = friction factor for smooth pipes. These equations present fairly good
approximations over a wide range of NRe.
The most widely used correlation between NRe and f (Moody, 1944) is shown in
Fig. 8-2. For an excellent discussion of fundamentals of flow in pipes, the reader is
referred to Szilas (1975) and Craft et al. (1962, pp. 1-100).

Friction head loss in fittings and connections

The Crane Company, which conducted exhaustive tests to find the resistance of
valves and fittings to single-phase flow, classified all valves and fittings as follows:
( 1 ) Branching- tees, crosses, side-outlet elbows, etc.
(2) Reducing or expanding-swages, reducers, chokes, bushings, etc.
( 3 ) Deflecting-elbows, bends, return bends, etc.
Crane Company also included an equivalent length concept, i.e., expressing head
loss due to the friction in valves and fittings in terms of equivalent head loss due to
the friction in a straight pipe.

In 1911, Blasius formulated the following empirical equation for turbulent flow:

f = 0.316/N;<'.

This equation is valid for smooth pipes at Reynolds numbers up to about lo5.
Formula 8-8 is commonly called Prandtl's formula and is expressed as follows:
Fig. 8-2. Friction factors for various types of commercial pipes. (From Moody. 1944 p. 671; courtesy of the Am. Soc. Mech. Eng.) NRr= V D p / p . where
V = velocity in ft/sec, D = pipe diameter in ft, p = fluid density, Ib/cu ft. and p = viscosity of fluid in Ib/ft-sec (cP/1488).

215
216

I'( - ")* FEET OF FLUID

SEE ALSO EQUATION (5)


IF A, = 00 SO THAT V, = 0
h = - 'I' FEET OF FLUID

V'
h = K - FEET OF FLUID
29

Fig. 8-3. Resistance coefficients for valves and fittings (reprinted from Engineering Data Book, 1979, p
75, table 32 (a); courtesy of the Hydraulic Institute, Cleveland, OH).
FI
I
217

REGULAR
-
BELL MOUTH
INLET OR REDUCER
SCREWED
45O ELL.
K
K=0.05

SQUARE EDGED INLET


K = 0.5

INWARD PROJECTING PIPE


K = 1.0

I
NOTE: K DECREASES WITH
INCREASING WALL THICKNESS OF
PIPE AND ROUNDING OF EDGES

LINE
FLOW
1

3'. .5 I n 2 A SCREWED
TEE

BRANCH
FLOW

h=K -
V'
2a
FEET OF FLUID
Fig. 8-3 continued.
21 8

0.wm
0.w15
0.w10
D
-
O.OW5
SMOOTH

0 1 2 3 4 s (I 7 8 8 10
R
-
D

Fig. 8-4. Resistance coefficients for 90" bends of uniform diameter. (From Engineering Data Book, 1979,
p. 79; courtesy of the Hydraulic Institute; Cleveland, OH).

The decrease in static head due to velocity is expressed as:


V2

and if there is a valve or fitting in the line, then the head loss due to the friction h , is

t
K
06

04

03

02

01

00
10 15 20 25 30 35 40

Fig. 8-5. Resistance coefficient for sudden reducers. (From Engineering Data Book, 1979, p. 82; courtesy
of the Hydraulic Institute. Cleveland, OH).
219

equal to:

V2
hl=K- (8-12)
2g
where K = resistance coefficient, defined as the number of velocity heads lost due to
the valve or fitting.
Inasmuch as the same head loss can be also expressed by the Fanning equation:

h 1 v2
=f--
d 2g

1
K=f- (8-13)
d
The ratio l / d is the equivalent length (in pipe diameters) of straight pipe, which will
cause the same pressure drop as the fitting under the same flow conditions. Thus,
the equivalent length (in ft) of a straight pipe, I,, is equal to:

(8-14)

where ( I / d ) = equivalent l / d for a given fitting, and d = diameter of the pipe in


which the fi&g is installed.
The K values for various valves and pipe fittings are presented in Figs. 8-3, 8-4,
8-5, 8-6.

PRINCIPLES OF PUMPING

Pumping mechanisms

The six basic mechanisms of artificially-induced fluid flow are (Perry and
Chilton, 1973): (1) the action of centrifugal force, (2) volumetric displacement
(mechanical or by another fluid), (3) mechanical impulse, (4)transfer of momentum
by another fluid, (5) electromagnetic force, and (6) gravity. The mechanisms (1) and
(2) are commonly used in the petroleum industry.

Measurement of performance

The amount of useful work performed by a fluid-transport device is the product


of the rate (capacity) at which fluid is transmitted and the head (the height of a
column of fluid equivalent to the pressure differential between inlet and outlet ends
of the device). Capacity can be expressed in ft3/min, whereas head can be expressed
in ft.
220

h
NDTF:

GL GATE VALVE -3,000


'14 CLOSED
'/2 CLOSED
- 2,000
' -FULLY OPEN
--~1,000 50
-
ANGLE VALVE, OPEN
I
STANDARD TEE -
7 500

- 300 30 30
- 200
22
/SQUARE 20
!
I-

F++
=loo
16
SWING CHECK VALVE, ,BORDA ENTRANCE 14
FULLY OPEN

--
-I

-
-- 0.5
- 0.3 I

- 0.2
LONG SWEEP ELBOW OR'
RUN OF STANDARD TEE
- 0.1
0.5

Fig. 8-6. Resistance of valves and fittings to flow of fluids. (Reprinted from Technical Paper No. 409,
1942; courtesy of Crane Co., New York).
221

The overall efficiency of the pumping system is defined as the ratio of useful
hydraulic work performed to the input work to the device.

8.33HGQ -
--H p
Hydraulic horsepower, P h = (8-15)
33000 1714
where H = total dynamic head, ft of liquid; H p = total dynamic head, psi; Q =
volumetric rate of flow, gal/min; and G = specific gravity with respect to water
( G , = 1). The brake horsepower of a pump, P b , is greater than the theoretical or
hydraulic horsepower by the amount of losses in the pump due to friction, leakage,
etc. The efficiency of the pump, vp, can, therefore, be defined as:

'h
v, =- (8-16)
'b

The total dynamic head H is equal to the total discharge head h , minus the total
suction head h , . The total suction head, h,, is equal to:

h , = h,, + ha + h,, (8-17)

where h,, = potential energy head at suction end, i.e., pressure head recorded on a
pressure gauge at the suction end in ft of liquid; h,, = velocity head, ft of liquid, at
the point of gauge attachment; and h a = atmospheric pressure head, ft of liquid.
Total discharge head, h,, is equal to:
h, = h,, + h a + h,, (8-18)

where h,, = pressure gauge reading at discharge end, ft of liquid, and h,, = velocity
head at discharge end, f t of liquid. Velocity head ( V 2 / 2 g )is equal to the vertical
distance necessary for a body to fall to acquire the velocity V.
There are limitations on the net positive suction head of the pump ( N P S H ) . The
maximum theoretical vertical distance between the pump and the level of suction
exposed to atmosphere is given by the following formula:

N P S H = h,, - h , - h , (8-19)

where h,, = static suction head, which is equal to the absolute pressure at the source
(ft of liquid) plus the vertical distance from this level to the pump center line;
h,, = friction head; and h , = head equivalent to the vapor pressure of the liquid.

OIL PIPELINE TRANSPORTATION

The flow of crude oil in a pipeline may be assumed to be isothermal if the oil
viscosity is low and the inflow temperature is close to the surrounding soil
temperature.
222

The design of a pipeline for hilly terrains has been discussed in detail by Beggs
and Brill (1972) and several others (see Chapter 9). Assuming a Newtonian liquid,
the principles are presented here.
As mentioned previously, the friction loss over a length of uniform pipe is given
by eq. 8-4:

This equation can be presented as follows:

(8-20)

where:

(8-21)

and h h is the hydraulic gradient.


Thus:

h , , =ALP = h I (8-22)
h
Y

where h,, = the total suction head at any point x along the length of the pipe and
y = specific weight of fluid.
One can then use the criterion of hydraulic gradient to determine the possible flow
over a given terrain, as shown in Fig. 8-7. Gradient h , is calculated using eq. 8-21,
and a line with a slope equal to this gradient is drawn from the outlet end 0 to the
inlet end. If this line A’ intersects the ground profile, there are two options:
(1) Increase the inlet head to h , . In this case, A’ is shifted upwards until it just
touches the profile at M. Because of safety considerations (safety factor), h , must
be higher than this minimum by 30-50 m.
(2) Increase the head h , to hma. This results in a line B having a different slope.
The point M is called the critical point, where the pressure head of the flowing
liquid is the lowest. If no throttling is applied at the tail end 0, the oil arrives at
atmospheric pressure, whereas if throttling is applied, the pressure head at the outlet
is h , .
The point M‘ in the valley is also critical, because one must calculate the pressure
head h and limit it to the maximum allowable operating pressure of the pipeline
( P= b ’ g ) .
The maximum flow rate in the pipeline can be determined as follows:
(1) Calculate the maximum head, h,, corresponding to the maximum allowable
operating pressure of the pipeline.
223

i
hl
7
1
09

Distance, f ? - f'
Fig. 8-7. Pressure profile for a pipeline laid over a hilly terrain. (Modified after Szilas, 1975, p. 479, fig.
7.1-1; courtesy of Akademiai Kiad6, Budapest, Hungary).

( 2 ) Determine the constants a and b in the equation f = a N i : for the flow


regime anticipated and the pipe in question. These are usually given in the pipe
manuals.
( 3 ) Substitute the above equation for f in eq. 8-21:

Inasmuch as V = -and g = 32.2 ft/sec-',


d277/4

(8-24)

or

(8-25)

This equation can be used to determine the maximum Q for a maximum h,, if
the only pump station is at the head end.
224

Example 8-1

Determine the maximum throughput of a 30-in. API 5LX grade X46 pipe
(maximum allowable stress, Sa(,,) = 21,000 psi), if the safety factor, F, = 1.2,
relative roughness c / d = l o p 6 (smooth pipe), d o = 30 in., d i = 28 in., Go,= 0.84,
poi,= 4 cP, pipe length = 40 miles, and outlet is located 400 ft above inlet.

Solution:

S, = Sam,/< = 21,000/1.2 = 17,500 psi.

Maximum allowable working pressure is given by eq. 8-3:

For additional safety, t h s h,, is, reduced by about 5%:

h,, = 3205 (0.95) = 3045 ft

Then, hh(,=) = (3045 - 400)/(40)(5280) = 0.012524 ft/ft

p, =4 CP= 4 x 6.72 X lb,/ft-sec =4 x 2.09 x l o T 5lb,sec/ft2,

p (4)(2.09)(10-5)
.. . v = o = = 5.2 x l o p 5 ft2/sec.
O po (0.84)(62.4)/32.2

Using a = 0.05 and b = 0.19 in eq. 8-24, Q can be calculated:

= 50.225 ft3/sec = 32,200 bbl/hr.

Increasing flow capacity of pipelines

There are two ways of increasing the flow capacity of an existing pipeline, which
should be achteved with a least-cost objective in mind:
(1) By installing one or more intermediate pumping stations along the pipeline.
(2) By installing a second pipeline alongside the already existing pipeline. This
so-called looped system is of two kinds:
225

(a) Complete loop-installation of a new line of the same length, but not
necessarily of the same diameter.
(b) Partial loop-installation of a new line shorter than the old line but of the
same inside diameter. This new line is preferably started at the tail end, in order to
achieve a lower average pressure in the entire system.
For short sections of the Moody friction factor chart, one can express f as an
exponential function of NRe:

f = aNi:

where a and b are constants. Assuming a = 0.05 and b = 0.19,

f = 0.05Ni:.19 (8-26)

From eq. 8-21:

2 f V 2 = 2(0.05)N;f,0.'9V2
hh= -
gdi 32.2di

- 0.05V2 diV
--
16.1di
-
(v)
~ 1 . 8 v0.19
1
= 0.0031 (8-27)
d;.l9

Inasmuch as V = Q / ( n / 4 ) d ? , in terms of Q, h , is equal to:

h h = 0 . 0 0 3 9 5 ~ ~-
"~( :*;:::) (8-28)

Thus:

Ql.81 = 252.95hhd,!.81
v0.19

or

(8-29)

and

(8-30)
226

These equations for h,, Q , and di can be used to calculate the parameters of the
new pipeline required.

( I ) Complete loop
Complete loop is regarded as two independent pipelines. In order to calculate the
diameter of the new pipeline, the following procedure must be followed:
If Qlmax=maximum flow rate in old pipeline (line l), Q, = total flow rate
desired, and Q, = flow rate in new Line 2 only, then Q , = Q, - elmax.
It is assumed that hhmaxis the same for both pipelines 1 and 2.
Equation 8-30 can be used to get the diameter of the new line d , (Q, = Q , , and
h h = hhmax):

di, = 0.3165( ~ ~ . ~ ~ ~ ~ Q 2 0 . ~ ~ ~ ) / h 2 ~ (8-31)

(2) Partial loop


In the case of partial loop for delivering at a rate Q, > elmax,
the gradient in each
line would be h , > hhmax.Using eq. 8-28:

hhl = 0.00395(

Inasmuch as the two pipelines have equal diameter and the same inlet pressure,
each will carry oil at a rate of Q,/2. Thus:

h,, = 0.00395( Q,/2)1'81/d4.81

Dividing hhl by h,, and solving for h,:

h,, = h h l ( i ) 1 . 8=
1 0.285hh1 (8-32)

After determining the two gradients, the length of the new pipeline can be either
calculated or determined graphically.
Graphical method. If the new pipeline is started at the head end Z, which is
done when pressure is to be maintained in the line due to hilly terrain, trace a
gradient h,, from the head end I and a gradient h,, from the tail end 0 as shown
in Fig. 8-8. The intersection point K enables determination of the length of the new
pipeline.
If the new pipeline is to be started from the tail end, which is preferable, then the
two gradients are reversed, as shown by the dashed lines meeting at point K'.
Calculation method. Refemng to Fig. 8-8:

h = h h m a x l = h h l ( l - l x+) h h , / ,
227

Dis?ance(l), miles W
-

Fig. 8-8. Pressure profile for a twin looped pipeline over a hilly terrain. (Modified after Szilas, 1975, p.
482, fig. 7.1-2; courtesy of Akademiai Kiad6, Budapest, Hungary),

Thus:

(8-33)

Example 8-2. Propose a feasible design for increasing the capacity to 75 ft3/sec,
using the data in Example 8-1.

Solution:
(1) For a complete loop:

Q , = 75 - 50.87 = 24.13 ft3/sec.

Using eq. 8-31:

d , , = (0.3165v0.0395Q,0~376)/h gz
- (0.3165) [ (5.128)(10 - )] 0.0395( 24.13)0.376
-
( 0.012784)0'208

= 1.756 ft = 21 inches.
The closest standard pipe size is Schedule 80, 24 in. OD, 21.562 in. ID, with a
wall thickness of 1.219 in.
228

(2) For a partial loop system:


Using eq. 8-28:

h hl = (0.00395v0.19Q:.81)/d4.81

= ((0.00395) [ (5 .128)(10-5)]0.19(75)'.8')/(28/12)4.8'
= 0.02543 ft/ft.

Using eq. 8-32:

h h , = 0.285hh, = (0.285)(0.02543) = 0.007248 ft/ft.

The length of the new pipeline in miles can be determined using eq. 8-33:

[x=
0.02543 - 0.012784
[ 0.02543 - 0.007248
]
(40) = 27.82 miles.

Booster pump stations


In designing booster pump stations, a simple graphical method can be used to
determine the minimum number of stations required.
Assuming that the oil is injected into the pipeline at pump stations at the
maximum allowable pressure of the pipeline, input pressure is the same at all
stations, except (usually) the last one. In designing for a minimum number of
pumping stations, the pressure gradient in the pipeline must reach the ground zero
level at the next pumping station along the line.
As shown in Fig. 8-9, starting from h,, at I , the gradient hh is plotted until it
intersects the ground profile. At this point, the second pumping station (the first
booster pump station) must be installed. Again the pressure at point B reaches h,,
and the same procedure is repeated for all the intersections with the ground profile,
until the outlet end 0 is reached. In Fig. 8-9 the last booster station will deliver oil
beyond point 0 (see Szilas, 1975, pp. 484-485).

Branching pipelines
Often, a single pipeline (called the main line or trunk line) is employed to
transport oil over large distances to a central location, from where it branches out to
various processing units. In order to determine the unthrottled throughputs of the
branch lines for any flow rate Q in the main line, both calculation and graphical
techniques can be employed (Szilas, 1975). It is much easier, however, to use a
simple log-log plot. The procedure is outlined below.
Consider a branched pipeline (Fig. 8-10.a) with flow rates Q , = 70 ft3/sec,
Q2 = 100 ft3/sec and Q3 = 150 ft3/sec. Then, using the h versus Q relationship for
these lines, draw the three h versus Q lines for the three branching pipelines. These
229

-
0

0'

Distonce(l), miles
Fig. 8-9. Pressure traverse with booster stations. (Modified after Szilas, 1975, p. 485, fig. 7.1-3; courtesy of
Akadtmiai Kiadb, Budapest, Hungary).

will all be parallel and are equivalent to a single line with a throughput.of Q,, which
is the sum of Q,, Q2 and Q3 (see Fig. 8-10.b).
From eq. 8-24:

3 'I 10 100 100


FLOW RATE, ~ ( f t J / r e c )
O3 @ G
Fig. 8-10. (a) Branching pipelines. (b) Flow rate versus head relation for branching pipelines. (Modifie
after Szilas, 1975, p. 487, fig. 7.1-5; courtesy of AkadCmiai Kiadb, Budapest, Hungary).
230

or

h, =KQ2-b (8-34)

where

2avb
K= is constant.
32.245 - b ( ;)2-

For a horizontal terrain, head loss h , is equal to discharge head h at the junction
B:

or log h = log K + ( 2 - b ) log Q . (8-35)

Equation 8-35 indicates that on a log-log paper, Q versus h relation is a straight


line. Inasmuch as ( 2 - b ) is a constant, the Q versus h relations for the branching
pipelines are parallel to each other, but have different K values.
The Q versus h relations for different pipelines are plotted on the log-log paper.
Adding the flow rates Q,, Q 2 ,. . . Q , of these pipelines yields a parallel line of the
same slope. This line is called the equivalent line and gives the sum of the
throughputs of the branch pipelines for any given head h. Thus, for any given total
flow rate Q , , one can get the throughputs Q,, Q 2 , ... Q , for the branch pipelines.
This is illustrated in Fig. 8-10.

NONISOTHERMAL FLOW

Pipelines transporting oil are commonly buried in the ground. If the oil viscosity
is high and the temperature of the flowing oil is significantly different from that of
the surrounding medium, the flow cannot be regarded as isothermal.

Fundamentals of heat transfer

Heat transmission can occur in three different ways: (1) conduction; (2) convec-
tion: (a) free and (b) forced; and (3) radiation.
When two bodies at different temperatures are in contact, conduction is the
dominant mechanism, whereas if the two bodies are not in contact but are separated
by a fluid body, the dominant mechanism is convection. In the latter case, the
temperature is transferred by the fluid in motion. If the two bodies are separated by
a vacuum, the mechanism of heat transport is radiation. Only conduction and
convection, however, are discussed in this chapter.
231

Conduction
Fourier's law is the fundamental differential equation for heat transfer by
conduction:
_
dq - dT
- - kA- (8-36)
dt dx
where dq/dt is the rate of flow of heat (quantity per unit time), Btu/hr; A is the
cross-sectional area across which heat flows, ft2; dT/dx is the rate of change of
temperature with distance in the direction of flow of heat (the temperature gradient),
"F/ft; and k is the constant called thermal conductivity, which is a characteristic
property of the material through which the heat is flowing and varies with
temperature, Btu/ft-hr-OF. Equation 8-36 can be used to derive the following
unsteady state energy equation for static fluids or solids in 3 dimensions:

cp( g) = &(kg) + $(k g ) + A(


) :k + q' (8-37)

where x, y , z are the distances along the x, y , z axes in Cartesian coordinates;


c = specific heat, Btu/lb,"F; p = fluid density, lb,/ft3; and q' = rate of heat
generation by chemical or nuclear reaction, electric current, etc.
In order to conform more closely to the physical shape of the system, this
equation can be transformed into cylindrical or spherical coordinates. In vector
notation, it may be written as follows:

(8-38)

Thermal conductivity, k, is approximately constant over a small range of tempera-


ture. Thus, eq. 8-38 can be simplified:
aT
CP-
at
= kV2T + 4' (8-39)

Steady-state conduction. For steady-state heat flow, dq/dt is constant and


aT/at = 0. Thus, eq. 8-39 becomes:
V2T= - q ' / k (8-40)
and eq. 8-36 may be written as follows:
dq/dt= -kA(dT/dx) (8-41)
where dq/dt = q / t = constant.
Unsteady stute conduction. In the case of unsteady state conduction, tempera-
ture is a function of both time and space. Equation 8-37 gives a general 3-dimen-
sional equation for such a situation. There is a numerical solution to this equation,
necessitating the use of digital computers. A variety of simplifications are incorpo-
232

rated to arrive at analytic solutions. One such case is the assumption of constant
physical properties c, p, and k , resulting in eq. 8-42:
aT 4'
at = a( V2T) + -
- (8-42)
CP

where (Y =k/pc, called thermal diffusivity.


Convection
Convection is an important factor in many cases of heat transfer involving
liquids and gases. When a fluid flows past a solid surface, in the immediate
neighborhood of the surface there is a film of fluid that does not contribute (or very
little) to the total flow. This film clings to the surface as shown by the velocity
distribution plot in Fig. 8-1. The film is in the laminar portion of the flow (the
laminar sublayer), through which heat is transferred by molecular conduction. The
turbulent core and the buffer layer between the laminar sublayer and the turbulent
core transfer heat by simultaneous conduction and convection. If physical proper-
ties are assumed to be constant, an energy balance equation on a flowing fluid
through which heat is being transferred is as follows:

(8-43)

where V,, 5,
V, are the velocity components in the x, y , and z directions and @ is
the energy dissipation due to the fluid viscosity, which is significant in the case of
flow of highly-viscous liquids and high-speed gas.
The local heat transfer coefficients were devised because of the impracticability
of measuring thicknesses of the several layers and their temperatures in a turbulent
flowing stream:
dq = hidAi(Ti - T,)
- = h , dAo(T, - To) (8-44)
dt
where h i and h , are the local heat transfer coefficients inside and outside the wall,
respectively (Btu/ft2-hr-"F), dA, = area element on the inside pipe wall; dAo = area
element on the outside pipe wall across which heat transfer is taking place, ft';
T, = wall temperature, OF; To = outside temperature, O F ; and T, = inside tempera-
ture, OF.

Natural convection
Natural convection occurs when a solid surface is in contact with a stagnant fluid
having a temperature different from that of the surface. The general equation used
is the Nusselt equation:
(8-45)
where NNu= Nusselt number, hI/k, (dimensionless); NGr = Grashof number,
233

I3p2gfiAT / p 2 (dimensionless); NPr= Prandtl number, c p / k (dimensionless); fi =


volumetric coefficient of thermal expansion, ( O F ) - ’ ; p = fluid density, lb,/ft3;
A T = bulk temperature gradient, O F ; 1 = length in contact, ft; g = acceleration due
to gravity, 4.18 x lo8 ft/hr2, and a and m are constants.
For cylinders at 1< NPr< 40, Kato et al. (1968) presented the following rela-
tions:

NNu= 0 . 1 3 8 N p (N:;17’ - 0.55) (8-46)

where NGr> lo9,

and NNu= 0.683N2~sN:;2s (8-47)


0.861 + N,,

where NGrG lo9

Forced convection
In the case of forced convection, the fluid is pumped across the solid surface and
the rate of heat transfer is a function of the physical properties of-the fluids, the
flow rates, and the geometry of the system.
For laminar flow in circular pipes, there are several relationships depending upon
the Graetz number, NGz= NReNp,d/f.For Ncz < 100, Hansen (1960) presented the
following relation:

( NNu)= 3.66 +
0.085Ncz h)’.l4( (8-48)
1 + 0.047NZ3 P w

where d = pipe diameter, 1 = pipe length, p b = viscosity of fluid at the bulk fluid
temperature, and p w = viscosity of the fluid at the pipe wall temperature.
For NGz> 100, the Sieder-Tate relation is satisfactory:
0.14
NNu= 1.86Ng3 (8-49)

For the transition region, the following equation was presented by Hansen (1960):

(8-50)

For turbulent flow ( N , , > 10,000) and 0.7 < NPr< 700, and f/d > 60, the Sieder-Tate
equation is recommended:
0.14
NNu= (8-51)
234

Application of heat transfer concepts to buried pipelines

Thermal properties of the soil must be evaluated when designing buried pipelines.
Heat transfer through the soil is proportional to its thermal conductivity, k, under
steady state conditions, whereas under transient conditions, the thermal diffusivity
a = k/cp is the controlling factor. It is, therefore, necessary to measure these soil
constants, k, p, and c, which can be done in the following two ways:

( I ) Estimation of thermal constants from soil properties


Thermal conductivity of the soil depends on the conductivity of the soil matrix,
grain-size distribution, bulk density of the dry soil, and humidity.
Gemant’s pioneering work was followed up by the studies of Makowski and
Mochlinski (1956) who derived an empirical expression for thermal conductivity of
wet soil in Btu/hr-ft-OF:

k, = 0.578( a log m + b)1O2 (8-52)

where a = 0.1424 - 0.000465(%~1);b = 0.0419 - 0.000313(%~1);%cl= percent of


clay in the soil; z = 0 . 0 1 ~ ~pd; = density of dry soil, lb,/ft3; and m = moisture
content of soil as a percentage of dry soil weight, 96.
In order to determine the thermal diffusivity it is necessary to calculate density of
wet soil, p,, in lb,/cu ft and specific heat of wet soil, c,, in Btu/lb,-OF. In terms
of pd and cd, these variables are equal to:

p, = pd(1 + 0.000624m) (8-53)

and

c, = (cd + 0.01m)/(1+ 0.01m). (8-54)

Combining the above three equations:

k, - 0.578( a log,,m + b)lOz(l + 0.01m) (8-55)


a,=--P W C W pd(l + o.o00624m)(Cd -k 0.01m)

(2) Estimation of thermal constants by direct measurement


Thermal conductivity is usually measured in situ by a field conductivity probe.
Heat is applied at a constant rate to this thin cylindrical probe which is pushed into
the ground. The rise in temperature of the probe surface is measured as a function
of time. The k value can be determined from the heat input and temperature-time
relationship.
Yearly average diffusivity of the soil can be obtained from the measurements of
the soil temperature at various depths below the surface. If the annual temperature
235

T i m e , t, sec +

u Tsmin -Pt
Reference Temperature (O'F)

Fig. 8-11. Fluctuation of soil temperature with time. Z = ground surface, ZZ = at a depth x . Amplitude
decreases with depth. (Modified after Szilas, 1975, p. 493, fig. 7.2-1.)

cycle is assumed to be sinusoidal (see Fig. 8-11), then the temperature change on the
ground surface, AT,, is equal to:

AT, =AT,,, sin at (8-56)

where A denotes temperature fluctuation and w = 277/t,, with t, = time for 1 cycle.
Figure 8-11 shows how the amplitude decreases with depth and that the tempera-
ture wave I1 at depth x is displaced in phase with respect to the ground-surface
wave (see Szilas, 1975, p. 493). The temperature variations can be expressed as:

aT a2T
--=ff- (8-57)
at ax2
where x = depth from the ground surface (x = 0 at ground surface).
Using eq. 8-56 in solving eq. 8-57,

AT, =AT,,, exp( -x/+) sin(21it/tp - x/-) (8- 58a)

and phase shift

At, = ( x / 2 ) { F (8-58b)

When the sine term in eq. 8-58a is equal to unity, the maximum temperature
fluctuation (amplitude) at a depth x is obtained:

AT,,, =A exp I- /x *s/at,


qmax (8-58~)
236

Example 8-3. What are the extremes of daily temperature and the phase shift
(a) at a depth of 1 ft and (b) at a depth of 5 f t if a = 5.3 X lo6 ft2/sec and the
surface temperature varies from 90 to 40°F during the day.

Solution: See Fig. 8-11.


The surface amplitude of the temperature fluctuation is equal to:

Using eq. 8-58a, the amplitude at a depth of 1 ft is equal to:

AT,,, = 25 exp[ - (l){r/(86,400 x 5.3 x ] = f1.82"F.


At a depth of 5 ft, the amplitude is equal to:

AT^,^ = f 5 . i x 10-SOF.
In order to evaluate the temperature extremes, first it is necessary to calculate the
mean temperature, which is approximately equal to:

Then, the temperature extremes at a depth of 1 ft are equal to:

T,,, = + AT,,, = 65 + 1.82 = 66.82"F


and
-
Tlmin= T, - AT,,, = 65 - 1.82 = 63.18"F.

At a depth of 5 ft:

= 65.00005"F and TSmin


TSmax = 64.99995"F.

By using eq. 8-58b, the phase shift at a depth of 1 ft is equal to:

At, = (;)/86,40O/(r X 5.3 X

= 36,017.5 sec = 10 hr.


Similarly, at a depth of 5 ft:

At, = 50 hr.
237

Steady-state flow in buried pipelines


Temperature of the oil injected into a pipeline usually differs from that of the
soil. Soil is generally cooler than the oil. Some of the factors contributing to the
variation of axial oil temperature in the pipe are:
(1) A part of the potential energy of the oil flowing in the pipeline is transformed
into heat, due to the shearing at the liquid-solid pipe wall interface and the
liquid-liquid shear within the bulk liquid. This heat increases the oil temperature.
(2) Temperature also increases due to the exothermic process of separation of
solid components (such as waxes) from the oil. Inasmuch as extensive separation
leads to reduced flow efficiency and eventual blocking of the oil, crude oils are often
dewaxed before pipeline transportation.
(3) Temperature decreases due to the transfer of heat from the pipeline to the
lower-temperature soil.
In calculating the heat generated by friction, one can assume an average constant
friction gradient along the pipeline. A pressure differential, Aprriction, is equivalent
to a force F, = A p , (7rd2/4) in lb, along the length I of pipe. (1 ft-lb, = 0.001285
Btu). If the flow rate is Q (ft3/sec) and the heat generated by F, is q, then velocity,
V, is equal to:

V = Q/( 7rd2/4)

and q = 0.001285 F,I.


The heat generated by friction per unit time per unit distance of flow, @J

(Btu/sec-ft), is equal to:

@ = 4 =0.001285FrI/t = [0.001285(A p , )( 7rd2/4) V ]/ I = 0.001285AprQ / l


It I
(8-59)

Using constant w and A, which are actually temperature dependent, the heat
liberated per one O F drop in temperature (Btu/'F-sec) is equal to:

where X = latent heat of fusion of the solids (paraffins) that separate out of the oil,

Y Surface

Fig. 8-12. Schematic diagram of an infinitesimal element of flowing fluid, exhibiting a temperature drop
dq.
238

Btu/lb,; w = amount of solids formed per lb, of oil per O F drop in temperature,
lb/lb,il-oF; p = density of the paraffin-containing oil, lb,/ft3; and Q = oil flow
rate, ft3/sec.
On considering an infinitesimal volume element, as shown in Fig. 8-12 where the
temperature of the flowing oil decreases by dT, over a length d l of pipe, one can
obtain, by heat balance:

PQC, dT, = @ dl + QpwXdT, - k(T, - T,) d l

Change in Heat Heat Heat lost


heat content generated generated to
of the by by separation surroundings
flowing friction of paraffins
liquid

where T, = axial temperature of the flowing oil in the pipeline, OF; T, = original soil
temperature at the same depth, OF; k = heat transfer coefficient per unit length of
pipe, Btu/ft-hr-OF; and c, = specific heat of the flowing oil, Btu/lb,-OF:
Or :

(Qpc,-AQpw) dT,=(@+kT,-kT,) d l (8-61)

On substituting Qpc, - QpwX = A (constant) and Qz + kT, = B (constant) in eq.


8-61, integrating (initial conditions: I = 0 and T = TI):

/ ( B - kA
T,) dT,=/dl

Solving, and resubstituting for A and B, one obtains:

(8-62)

The oil temperature also varies radially. In turbulent flow, the radial temperature
variation is slight, whereas in laminar flow, it may be significant. In the latter case,
the axial temperature Tfis somewhat higher than the average temperature. Exam-
ples of temperature profiles are shown in Fig. 8-13. In practical calculations, the
latent heat of fusion of paraffin is usually not considered. As shown in Example 8-4,
however, it may be significant.

Example 8-4
Oil is injected at a flow rate Q = 700 bbl/hr into a 20-mile long pipeline having
I.D. = 12 in. The oil gravity is 0.85, temperature T, = llO°F, viscosity is 4.2 CP
239

Velocity + Temperoturr

(a) I b)

Fig. 8-13. Velocity (a) and temperature (b) profiles for oil flowing in a pipeline. (Modified after
r=
Chernikin, 1958 in: Szilas, 1975, fig. 7.2-4, p. 498). r=
0.317 m/sec, 69"C, T, = 55.5"C.

(8.778 X slugs/ft-sec), soil temperature T, = 40 OF, relative roughness c/d =


0.0001, w = 0.003/"F, c, = 0.455 Btu/lb-OF, X = 200 Btu/lb,, and k = 1.2 Btu/hr-
ft-OF. There are 2 globe valves, 1 long sweep elbow, and 100 couplings present. Find
the exit temperature Tn of the oil:
(1) Assuming friction and paraffin deposition over the entire length of the pipe.
(2) Assuming negligible paraffin deposition.
(3) Neglecting the effects of both friction and paraffin deposition.

Solution: Velocity of oil in the pipe, V = Q/( 7r/4)dz = [700 X 5.615/ 3600
ft3/sec]/[(7r/4)(1)2 ft2] = 1.39 ft/sec. Oil specific weight, y = 0.85 X 62.4 = 53.04
lb/ft3, whereas oil density, p = 53.04/32.2 = 1.647 slugs/ft3( = 53.04 lb,/ft3).

N Re = -d=Vp (1)(1.39) (1.647)


= 26,100.
P (8.778 X

From Fig. 8-2, for relative roughness r/d = 0.0001 and NRe= 26,100, the Fan-
ning friction factor f = 0.0242. From Fig. 8-6, I, for a globe valve = 320 ft, and I,,
for a long sweep elbow = 20 ft. From Fig. 8-3, K for a coupling = 0.04. Thus, the
head loss due to friction is equal to:

2f/V2y V2y 2(0.0242)[(20)(5280) + 2(320) + 201 (1.39)2(53.04)


*Pf= gdi+ K - 2g =
(32.2) (11

Using eq. 8-59:


240

(1) From eq. 8-62, the exit temperature is equal to:

110 - 40 - -
0.783 x exp[
1.2
1 - (1.2)( 20)( 5280)
(700 X 5.615)(53.04)(0.455 + 0.6) 1
+--0.783
1.2
-79.63'F

(2) On assuming w = 0,

0*783
110 - 40 - -
- (1.2)(20)(5280)
1.2 x exp[ (700 X 5.615)(53.04)(0.455) 1
+-
0.783
1.2
=58.88"F

(3) On assuming w = 0, and @ = 0,

T,, = 40 + (110 - 40) exp


[ - (1.2)(20)(5280)
(700 X 5.615)(53.04)(0.455) 1 = 58.40"F

This example enables one to evaluate the significance of each of these assump-
tions. The paraffin deposition, which is neglected by many pipeline engineers, has a
very significant effect.
It is important to understand what the heat transfer coefficient, k, per unit pipe
length signifies. As shown in Fig. 8-14, there are a series of resistances to the
transfer of heat from the oil to the surrounding soil.
First there is heat transfer from the bulk oil to the pipe wall, which is a process of
forced internal convection (see eqs. 8-48, 8-49, 8-50, and 8-51).
For Na -= 100 and NRe < 2100 (eq. 8-48):

NNu=3.66 [ 0.085NGz
+ 1 + 0.047NZ3 ]( e) 0.14

Fig 8-14 Cross-sectional view of a buried insulated pipeline carrying oil


241

For NGz > 100 and NRe < 2100 (eq. 8-49):

NNu= 1. 8 6 N g 3
i-:L r4
For 2100 < NRe < 10,000 (i.e., the transition region; eq. 8-50):

(”)
0.14
”, = 0.116( N i l 3 - 125) N ; i 3 [ 1+ ( d / ~ ~ / ~ ]
PW

For NRe > 10,000 (turbulent flow; eq. 8-51):

0.14
N,, =

These equations correspond to heat transfer by convection in the three zones of


flow, i.e., laminar, transition, and turbulent. Thus, one can evaluate the Nusselt
number N N u using the appropriate equation for a particular flow regime.
Inasmuch as:

hidi
NNu= - (8-63)
k

k
h i = NNuX - (8-64)
di

In this manner, the heat transfer coefficient h (Btu/hr-ft2-OF) can be de-


termined. It is called h i , because it is the heat transfer coefficient inside the pipe.
Thus, the net heat transfer per unit length of pipe (Btu/hr-ft-OF) due to the internal
forced convection is equal to:

Another resistance to heat transfer is the pipe wall. If the thermal conductivity of
the pipe wall material is k,, then its net contribution in Btu/hr-ft-”F is equal to
2rkw/ln( do/di).
In the presence of thermal insulation around the pipe, another factor
(2nkin/ln( din/d,) is introduced, where kin= thermal conductivity of the insulating
material and din= diameter of the insulating material.
Finally, the thermal conductivity of the soil surrounding the pipeline must be
considered. Assuming a constant undisturbed soil temperature T, at a depth x , it
242

can be shown that the net heat transfer coefficient of soil, h,, is approximately
equal to :

In order to obtain the net effective resistance, it is necessary to add these four
resistances ( = l/conductance) in series together:

R=- 1
hi(rdi) + -
1
2 r k w ln( 3)+ - 1
2rk,
In( 3)+ 1
-
hordin
(8-66)

Thus, the total heat transfer coefficient per unit length is equal to:

k- = - 1= r
(8-67)
‘ R
-+ +
1 1 1 1
- In(%) + - ln(2) -
hidi 2kw di . 2kin hodin

The kin values for some common insulating materials are presented in Table 8-1.
Pipes, which are usually laid in ditches dug for that purpose, are covered with
backfill. Inasmuch as the porosity of this backfill is greater than that of the
undisturbed soil, its thermal conductivity is lower. The backfill compacts with time,
however, causing the effective heat transfer coefficient to increase slightly. Other
variables are wind, plant cover, snow, moisture content of the soil, etc.
Under steady-state conditions, the heat transfer across the flowing oil into the
soil is equal to:

= 2rkin(T0- qn)/ln(dh/do)

= hodin..( qn- T , ) (8-68)

One can assume that the inside and outside wall temperatures of the pipe are

After Rohsenow and Hartnett (1973) who gave an appropiate relationship:

2ks
h, =
dincosh-’(2x/din)

where cosh-’(2x/din) is approximately equal to In(4x/din).


243

TABLE 8-1
Thermal conductivities of pipeline insulators (after Thomas and Turner, 1953; courtesy of Chemical
Engineer)

Material Resistance Temperature limits Thermal conductivity Specific


towater (OF) (Btu/hr-ft 2-oF/in.) weight
Min. Max. 32°F 70°F 212OF 500OF
Cellular glass Excellent -400 800 0.35 0.39 0.46 - 10
Cellular silica Excellent - 300 1600 - 0.44 0.62 1.0 10-12
Cement Water will 100 1800 - 0.69 0.820 1.058 26
(semi thermo- soften the
setting) dried cement
Cement Will not 32 1200 0.46 0.525 0.610 0.840 49
(hydraulic soften when
setting) wetted
Diatomaceous
earth Fair 22 1900 - 0.66 0.720 - 23
Burnt diatomaceous
earth bricks Moderate 22 1900 0.603 0.617 0.694 0.742 29.95
Burnt diatomaceous
earth paste Moderate 22 1900 0.520 0.541 0.590 0.610 31.2
Glass fibers Excellent - 300 600 0.270 0.273 0.320 - 8
(formed into
pipe insulation and
blocks)
Hydrous calcium
silicate Good 100 1200 - 0.37 0.41 0.52 12
Perlite Excellent 32 1200 - 0.38 0.43 0.58 10-12
Polystyrene Excellent - 300 175 0.23 0.26 - - 2-2.3
(expanded)
Polyurethane
(expanded) Good - 50 230 - 0.17 0.35 - 3

equal ( q= To),and then break up the above multiple equation into three indepen-
dent equations:

k ( T , - T,) =mdihi(T,- Ti) (8-69)

(8-70)

and

hidi&- Ti) =hodinpin- T , ) (8-71)

If Ti is unknown, one should use a trial and error procedure. Assume Ti and
establish the physical parameters of the oil (density, viscosity, specific heat, etc.) to
244

Fig. 8-15. Universal viscosity versus temperature chart for crude oils. (After Frick, 1962, p. 19-39, fig.
19-41; courtesy of McGraw-Hill, New York).

evaluate NRe, NP,, NGr,Na and NNu. Knowing NNu,calculate h i and then k
using eq. 8-67. Another k value is obtained using eq. 8-69.
If the k values obtained by the two methods agree, then the assumed value of Ti
is correct. If not, the calculation procedure is repeated with a different value for Ti,
until one arrives at an acceptable agreement.
To carry out such calculations, the variation of the physical parameters (viscosity,
density, etc.) of the oil with temperature must be known.

Viscosity
Temperature and liquid viscosity may be correlated within the accuracy of most
experimental data (1-2%) with the de Guzman-Andrade Equation:

where A and B are constants.


This equation implies that a plot of log p versus 1/T will yield a straight line.
Figure 8-15 illustrates the above equation, which is generally used.
Beggs and Robinson (1975) presented a more recent empirical correlation be-
tween viscosity and temperature, which gives better results than the frequently used
Beal's (1946) correlations:

pdo = 10" - 1 (8-73)

where pdo= viscosity of dead oil (gas-free crude oil), cP; and x = Y T - ' . ' ~ ~where
,
T - O F ; y = 10'; and z = 3.0324 - 0.02023 ("API).
245

Fig. 8-16. Viscosity of gas-free crude oil at reservoir temperatures. (After Beal, 1946, fig. 8, p. 103;
courtesy of the S.P.E. of A.I.M.E.).

I00
80
60

40

20

10
8
6

I
08
06

04

I-
4, 0 2

1111 I I I1 I IllllJ
0.4 OK08 I 2 4 6 8 10 20 40 60 BOIOO
VISCOSITY OF DEAD OIL, CENTIPOISES
(AT RESERVOIR TEMPERATURE AND ATMOSPHERIC PRESSURE)

Fig. 8-17. The viscosity of gas-saturated crude oils at reservoir temperature and bubble-point pressure.
(From Chew and Connally, 1959, fig. 2, p. 25; courtesy of the S.P.E. of A.I.M.E.).
246

Fig. 8-18. Relationship between the viscosity of oil and pressure (above the bubble-point). (From Bed,
1946, fig. 11, p. 109; courtesy of the S.P.E.of A.I.M.E.).

For crude oil containing dissolved gas, Beggs and Robinson provided the
following correction for viscosity:

Po =A d o (8-74)

+
where A = 10.715(RS+ 100)-0.515;B = 5.44 ( R , 150)-0.338;and R , = amount of
dissolved gas in oil, scf/STB.
This equation gives the viscosity of gas-containing (live) oil. This correlation was
developed by plotting log,, T versus log log(pdo+ 1). Straight lines were obtained
and it was found that each of these lines corresponded to crude oils having a
particular API gravity, which led to the development of eq. 8-73.
Because of their popularity, Beal’s correlations are also presented in Fig. 8-16 and
8-18, along with the Chew and Connally (1959) correlation (fig. 8-17).

Density
The following equation takes into account the effects of both temperature and
pressure on the density of the oil:

or

where p T = density of oil at any temperature T, lb,/ft3; pn = density of the oil at.
the base temperature T, and base pressure (usually at standard conditions); a T =
temperature coefficient of oil density, Ib,/ft’-’F; ap = pressure coefficient of oil
density, l/ft; p = pressure, lb/ft2; and y = specific weight, lb/ft3.
247

The coefficients aT and ap can be evaluated if the densities are known for any
two known conditions of temperature and pressure. There are other empirical
correlations which are not presented here.

Kinematic viscosity
A useful correlation between temperature and kinematic viscosity Y (which
combines both the viscosity p and the density p of the oil) was proposed by
Walther. This equation, however, neglects the effect of pressure, which is generally
small:

log log(102v + a ) = b + c log T (8-76)


where a , b, and c are constants, their value depending upon the particular oil in
question. The value of a is around 0.8 for most of the oils. On incorporating this
value into the above equation, one obtains:

log log(102v + 0.8) = b + c log T (8-77)


where Y = cm2/sec (or stokes) and T = temperature, "Rankine.

Specific heat
Cragoe (1929) presented a correlation for the estimation of the specific heat cp in
Btu/lb,-'F for petroleum oils:

c =
0.388 + 0.00045T (8-78)
P G,0.5
\

where T is the temperature in O F and Go is the liquid specific gravity at 60°F. The
accuracy of this equation is +5%.
Thermal conductivity
The American Petroleum Institute (1966) recommended a single value of 0.077
Btu/hr-ft-OF at 30°C (86°F) for thermal conductivity of petroleum oils. The
average and maximum deviations from this value are 7 and 308, respectively. At
other temperatures, Cragoe's equation gives satisfactory results, with average and
maximum deviations of 12 and 398, respectively:

0'0677 [ 1 - 0.0003( T - 32)]


k= - (8-79)
Go

where 32" < T -= 392°F.


Example 8-5. Determine the heat transfer coefficient for a 10-mile pipeline
buried 4 ft below the surface. Pipe having O.D. = 12.75 in. and I.D. = 12.438 in.
carries a crude oil having specific weight of 53 lb/ft3 (Go= 0.85 or 35'API). Given:
= 110'F; T, = 35'F; aT = 0.25 1b/ft2-'F; k , = 1.0 Btu/hr-ft-"F. Assume laminar
flow: NRe= 2000.
248

Solution: For the first trial, Ti = 80°F is assumed. Using eq. 8-73:

z = 3.0324 - 0.02023(35) = 2.32435

y = 10' = 102,32435= 211.033

x =yT-'.163 = (211.033)(80)-
1.163 -
- 1.2914

... = 101.2914- 1 = 18.56 CP

and

p,(at T,) = 6.8 cP.

Neglecting the effect of pressure, the density is equal to (eq. 8-75):

p r = p , - a T ( T - T,)=53.0-0.25(80-60)=48 lbm/ft3.

Using eq. 8-78, specific heat of oil is equal to:

The thermal conductivity of oil can be determined from eq. 8-79:

k o = -[l - 0.0003(80 - 32)] = 0.0785 Btu/hr-ft-"F


0.85

The Graetz number, N,(

NGz = (2000)
= N R e N P r d / l )is

0.0785
3600
Btu
sec-ft-OF
Ibm
equal to:

6)7X -
(18.56 X 6.72 X 1 0 - 4 ) ( 0 . 4 7 1Btu lbm
ft-sec 1 [lzO]
X
12.438
= 10.6

Inasmuch as Ncz < 100, one can use eq. 8-48 to find NNu:

NNu = 3.66 +

h i= NNux k o = (0.844)[
-
di
1i:;l;i2] = 0.0635 Btu/hr-ft2-OF
249

and

-
2)
h, = = 0.6941 Btu/hr-ft2-"F

)
4x4
din ln( In( 12.75/12

Thus, k can be determined using eq. 8-67:


1T
k= = 0.0604 Btu/hr-ft-'F
1/(0.06352 X 12.438/12) + 1/(0.6941 X 12.75/12)

and using eq. 8-69:

mdihi(T, - q )
k= = [ ~(12.438/12)(0.0635)(110 - 80)]/(110 - 35)
( T f - T,)

= 0.0827 Btu/hr-ft-"F

On assuming Ti = 82"F, as a further approximation, the above procedure is re-


peated :
Equation 8-67 gives k = 0.0711 Btu/hr-ft-"F, whereas eq. 8-69 gives k = 0.0705
Btu/hr-ft-OF, which is a reasonable agreement. Thus, k = 0.071 Btu/hr-ft-"F and
Ti = 82°F.

Transient (unsteady state) flow of oil in buried pipelines

If the inlet temperature, flow rate, the physical parameters of the oil, and the soil
temperature are constant over a comparatively long period of time, the heat flow in
and around a buried pipeline must be steady. These conditions are, however, never
satisfied. Slight departures from the steady-state behavior are approximated by the
steady-state relationships. In numerous practical situations listed below, however,
the departure may be sufficient to render the steady-state relationships useless for
even an approximate quantitative evaluation:
(1) Termination or beginning of flow in the pipeline.
(2) Change in injection temperature of oil into the pipeline.
(3) Fluctuations in temperatures.
(4) Fluctuations in flow rates.
Given sufficient time, any kind of flow (except the fluctuating type) will
eventually stabilize and reach steady state.
Numerous models and calculation methods have been devised to describe and
solve unsteady-state systems. Szilas (1975) presented a transient model which
describes the temperature changes in a pipeline shut down after steady-state flow.

' This book was originally published in Hungary in 1968


250

He called it a “cooling model” for an insulated pipeline.


In the case of shutdown after flow, it is assumed that:
(1) The pipeline is embedded in an infinite half-space filled with soil of homoge-
neous thermal properties.
(2) The temperature of the soil in contact with the pipeline can be described by
Chernikin’s (1958) model.
Heat flowing through the wall of unity pipe length into the cooler ground over an
infinitesimal period of time d t reduces the temperature of oil and pipe by dT:

[ rd?
4po~o
1
+ -774 ( d ? - df ) pPcp dTo = - k (T; - Tin) d t
where subscript o refers to oil, p refers to pipe, and in refers to pipe insulation; T,,
and T[n are transient temperatures. Neglecting h i and h , and using eq. 8-67:

k = r/ [ (1/2kin) In( din/do )I


Setting:
77 [ dfpoc0 + (d,2 - df)PpCp] = 4
-
4k
and rearranging:
A[dTo/(T,’- T6)] = -dt (8-80)
The Chernikin relationship is then used here to describe the change in the oil
temperature, TA, and the change in the outer temperature of the insulation Tin. If
point PI is the image of the projection of the linear heat source on a plane
perpendicular to it (see Fig. 8-19), then the difference in temperature between the
point Pz (lying in the plane of projection and defined by the coordinates y and z)
and the undisturbed soil is equal to:
TL- T,= (@/4ak,)[Ei(-xz/r2X1/NF,) -Ei(-1/4NFo)] (8-81)

where NFo= q / r 2 = Fourier number (dimensionless), and Ei is the exponential


integral function.

Fig. 8-19. The Chernikin model for unsteady state flow of oil in buried pipelines. (Modified after
Chernikin, 1958; in: Szilas, 1975, p. 517, fig. 7.2-13; courtesy of Akadkmiai Kiad6, Budapest, Hungary).
25 1

If t = 00 (steady state flow), eq. 8-81 reduces to the following form:

To - T, = (@/21rk,) ln(2x/r) (8-82)


Dividing eq. 8-81 by eq. 8-82:

Inasmuch as x, r , k,, ps and c, are constant for a given pipeline, the above equation
can be written as follows:

where

The process of warming 'up is described in eq. 8-85. In the case of cooling down
after steady-state flow:
( T; - T , ) / ( To- T,) = 1 - x
or
Td=(l-X)(To-T,)+T, (8-86)

The relationship X = f(t) for a given case can be plotted on a graph using eq.
8-86. Sections of this curve can be individually approximated by the following
relation:
X=a+bht (8-87)
In addition, an equation similar to eq. 8-86 can be written for the insulation around
the pipe:

TL = (1 - X ) ( Ti,- T,) + T, (8-88)

Combining eqs. 8-87 and 8-88:


Ti',= (1- a - b h t )( Ti,- T,) + T, = T, + (1- a ) ( Tin- T,) - b( Ti,- T,) In t
or
Ti:, = B - C In t (8-89)

where B = T, + (1 - a)(Ti, - T,) and C = b(Ti, - T,).


252

Substituting q,,from eq. 8-89 into eq. 8-80:

’[ T d - BdTo
+Clnt I = -dt

The general solution of eq. 8-90 is:


(8-90)

T,’=B-C In t + [CEi(t/A)C’] e-‘IA (8-91)


where C’ is the constant of integration.
In the case of initial condition T,’ = at t = 0:

Ti = B - C In t + [ CEi( t/A)To - B - C(0.5772 - In A ) ] e-r/A (8-92)


The above relationship can be used to evaluate the variation of temperature Ti
versus the time t elapsed after shutdown of the pipeline at any pipe section situated
at a distance 1 from the head end of the line (See Szilas, 1975, p. 519.).
Heating up of a cold line by introduction of hot oil
When hot oil is pumped into a cold line, the initial transient heat loss-to the cold
soil can be very much higher than the equilibrium steady-state heat loss. In large
crude oil lines, it may take many days or even weeks before equilibrium conditions
are established. It is, therefore, necessary to know how the transient heat loss varies
as a function of time. (See Davenport and Conti, 1971.)
In the simple case, the pipe is assumed to be surrounded by an “infinite sea” of
soil having homogeneous thermal properties. The heat transfer between the ground
surface and air is assumed to be infinite. Thus, the heat loss is directionally
symmetrical and for this simple case, the Nusselt number is related to the Fourier
number as follows (see Davenport and Conti, 1971):

N,, = 0.362 + 0.953/N;:’ (8-93)


The soil will warm up continuously and will reach equilibrium with the pipe
surface. It is assumed that equilibrium is reached when the Nusselt number from eq.
8-93 equals that from eq. 8-94:

(N,,), = 2/cosh-’(2x/d0) = 2/ln(4x/d0) if x >, do (8-94)

Thus, it is possible to calculate the time necessary to reach equilibrium, which is a


function of NFo and x / d o . Due to the approximation employed, NNuis a function
of x/do only in eq. 8-94. Substituting eq. 8-94 into eq. 8-93, yields:
r

7 (8-95)
253

where N,, = cy,t/d:; x = depth of burial of pipeline, ft; and do = outside pipe
diameter, ft.
This relationship gives the time of transient flow in the pipeline as a function of x
and do.
As the temperature wave from the pipeline reaches the surface and is reflected,
the surface begins to play an important role and the infinite sea approximation is
invalid. The heat loss exceeds the calculated value by 10 - 15%thereafter.

TRANSPORTATION OF HEAVY OILS IN PIPELINES

Heavy oils, characterized by high viscosities, high pour points, and low API
gravities, are currently being transported only to a limited extent by pipelines.
Development of the largely untapped world resources of heavy crude oils, however,
is changing this. The methods for pipelining heavy crude oils are briefly discussed
here. (See Barry, 1971.)
The major problems are caused by: (1) pour point (wax crystallization problems)
and ( 2 ) viscosity (flow problems). High pour point temperatures are caused by
excessive formation of wax'crystals in the oil, which inhibit its ability to flow. Wax
deposits form in the storage tanks as well as in the pipeline. With increasing
viscosity, the head loss due to friction increases and, therefore, greater pump
horsepower is required. Thus, it is necessary to reduce the viscosity and the pour
point of the oil being transported, which can be achieved by the following means:
(1) Use of additives (pour point depressants).
( 2 ) Preparation of an oil-solid slurry.
(3) Heating the oil to keep the flow essentially above the pour point, and also to
reduce viscosity.
(4) Dilution of oil with a solvent or with another oil for reduction of viscosity
and pour point.
( 5 ) Preparation of a lower-viscosity, unstable slurry-emulsion system by mixing
water with the oil.
The first two techniques are applicable only in cases where the oil viscosity is low
enough to permit economical pipelining at the existing temperature. The third and
fourth techniques are currently the major methods for the pipelining of heavy crude
oils. The fifth technique has not been used yet on a commercial scale.
The transportation of hot oil by pipelines, which has been discussed already,
requires some means of heating the oil. Usually direct-fired heaters are used. The
dilution technique involves the addition of low-viscosity hydrocarbons such as
condensate, natural gasoline, and naphtha in order to reduce viscosity of heavy oil.
In the case of oil fields located in remote areas, where the diluent may not be readily
available, two major alternatives are available: (1) Use of a dual pipeline, so that
one pipeline could carry the diluent back to the field. ( 2 ) Installation of a thermal
cracker, catalytic cracker, or a hydrocracker to crack a portion of the heavy crude to
254

lighter components, which can then be blended with the remaining heavy oil for
pipelining.
The choice of a technique is determined by the oil viscosity, the geographic
location, length of pipeline, process feasibility, and economic considerations.

PIPELINE TRANSPORTATION OF NATURAL GAS

Physical properties of gases

Physical properties of natural gases (compressibility, density, viscosity, and


specific heat), which affect gas transmission in pipelines are discussed first in this
section.

Gas compressibility
Deviation from ideal behavior of natural gas is seldom negligible. For practical
purposes, only compressibility factor Z is applied:

p v = ZnRT (8-96)

GAS GRAVITY (AIR=I)

Fig. 8-20. Pseudocritical properties of condensate well fluids and miscellaneous natural gases. (After
Brown et al., 1948; courtesy of Natural Gasoline Association of America).
255

where p = absolute pressure of the gas, psia; T/;= total gas volume, ft3; Z =
compressibility factor; n = number of moles of gas in volume y ; T = absolute gas
temperature, OR; and R = universal gas constant = 10.73 psia-ft3/lb mole-OF.

PSEUDO REDUCED PRESSURE PR


Fig. 8-21. Compressibility factor for natural gases as a function of reduced pressure and temperature.
(From Standing and Katz, 1942, fig. 2, p. 144; courtesy of the S.P.E.of A.I.M.E.).
256

According to the law of corresponding states, the compressibility factors Z of


two gases are equal if the reduced state parameters of these gases are equal. This is
the basis for the generalized compressibility factor chart shown in Fig. 8-21.
Reduced pressure p , is equal to p / p c and the reduced temperature T, is equal to
T/T,, where p, and T, are the critical pressure and temperature of the gas,
respectively. For a mixture of gases, the reduced state parameters are replaced by
the pseudo-reduced parameters ppr and Tp,, which are defined as follows:
n

~ p =c C Yipci (8-97)
i=l

and
n
Tpc = C YiTci (8-98)
i-1

where yi = mole fraction of component i in a gas mixture.


The molecular weight, M , is equal to:
n
M= CyiMj (8-99)
i=l

If the gas composition is not known, one can use approximate empirical
correlations (see Fig. 8-20) developed by Brown et al. (1948).
Natural gas often contains nonhydrocarbon gases such as N, and CO,. If the N,
content is less than 8% and that of CO, is less than 108, one can use the additive
rule to determine compressibility factor Z :

where ZHc is the compressibility factor for the pure hydrocarbon gas, which can be
determined as before. The compressibility charts for N, and CO, gases are given in
literature.

Density
The gas density, p , at a pressure p and temperature T can be obtained from eq.
8-96 as follows:

n / y =p/ZRT (8-100)

Inasmuch as number of moles n = mass/molecular weight = m / M , eq. 8-100 be-


comes:

m/K=pM/ZRT=p (8-101)
251
Fig. 8-22. Relationship between the viscosity of paraffin gases and molecular weight at a pressure of orre atmosphere and reservoir temperatures, with

.E3
corrections for nitrogen, carbon dioxide, and hydrogen sulfide. (From Carr et al., 1954, fig. 6, p. 268; courtesy of the S.P.E. of A.I.M.E.).
258

Fig. 8-23. Viscosity ratio, p / p l , as a function of pseudoreduced temperature and pseudoreduced


pressure. (From Carr et al., 1954 p. 267, fig. 4; courtesy of the S.P.E.of A.I.M.E.). p = viscosity of gas at
reduced temperature, T,, and reduced pressure, P,; p1= viscosity of gas at atmospheric pressure and at
temperature T,.

Gas gravity, Gg, is defined as the ratio of the density of the gas, pe, to the density
of air, pa, under standard conditions:

(8-102)

because Z = 1 at standard conditions.


Inasmuch as molecular weight of air is equal to 28.97, eq. 8-102 can be written as
follows:

Gg= M/28.97

Viscosity
Gas viscosity decreases with increasing molecular weight and increases with
increasing pressure and temperature. The pressure effect is the same as in liquids,
whereas the temperature effect is just the opposite.
The two most widely used correlations are those of Carr et al. (1954, see Figs.
8-22 and 8-23) and Lee et al. (1966). Beggs and Brill (1972), gave the following
modified form of Lee's equation:

1.1= A x exp( BpC) (8-103)

where p is the viscosity in cP, A = (9.4 + 0.02M)T1.s/(209 + 19M + T ) ; B = 3.5 t


259

Temperature, OF

Fig. 8-24.Viscosity of natural gases at atmospheric pressure. (After C a r et al., 1954, fig. 7, p. 269;
courtesy of the S.P.E.of A.I.M.E.). Data is based on N.B.S.-N.A.C.A. tables of thermal properties of
gases and research work by J.O. Hirschfelder, R.B. Bird and E.L.Spotz (1949),M. Trantz and K.G. Sorg
(1931),and A.O. Rankine and C.J. Smith (1921).

+
986/T 0.01M; C = 2.4 - 0.2.4; M = molecular weight of the gas; T is the temper-
ature in OR; and p is the density in g/cm3.
The viscosities of natural gases at atmospheric pressure are given in Fig. 8-24.

Specific heat
Specific heat is defined as the amount of heat required to raise the temperature of
a unit mass of a substance by one degree. It can be measured at a constant pressure,
cp, or at constant volume, co, resulting in two distinct specific heat values. For an
ideal gas, the difference between the two is equal to the gas constant R:

cP - C, =R (8-104)

The cp of a gas mixture can be calculated from the following formula:

c
n
cp= YiC,i (8-105)
i-1

Product of gas constant R o and molecular weight M is called Universal Gas Constant ( R= R o M )
and is equal to = 10.732 p~ia-ft~/Ibmole-~
R = 1544 ft-lbf/lbmole-o R = 1.986 Btu/lbmole-o R.
260

.W

TEMPERATURE DEtREfS FAHRENHEIT

Fig. 8-25. Specific heat, cp, of hydrocarbon vapors at a pressure of one atmosphere. (After Brown, 1945,
fig. 1, p. 66; courtesy of the S.P.E.of A.I.M.E.).

Brown (1945) has presented the molar specific heats at constant pressure for the
individual hydrocarbon gases (Fig. 8-25). Equation 8-105 can be used to determine
approximately the specific heat of a mixture of gases if the composition is known.
261

0.04 0.060.000.10 0.2 0.4 0.6 0.8 I 2 4


Reduced pressure pr=P/P,

Fig. 8-26. Generalized relationship of heat capacity differences, cp - c,,, to reduced pressure and reduced
temperature. (After Edmister, 1948, p. 613, in: Perry and Chilton, 1973, fig. 3-54, p. 3-238; courtesy of
Petroleum Refiner.)

The specific heat capacity ratio, c,/c,, for an ideal gas can be easily obtained
using eq. 8-104. For a real gas, CJC, ratio is equal to:

The ( c p - c u ) quantity, which is larger than R for a real gas, may be obtained from
the generalized chart in Fig. 8-26.

Gas flow fundamentals

Detailed treatment of gas flow fundamentals was presented by the Institute of


Gas Technology (1972). Some important equations are presented here.

Q b = 38.774-

QbGgpb
p: -pz - 0.0375GgA~p,Z,,/(TZ) Os
G,ITZf
x d 2s
1 (8-107)

NRe = 0.4775 - (8-108)


pdiTb
262

(8-109)
r 1

(8-110)

(8-111)

(8-112)

Pavg = 0.67( -) (8-113)

where Q,=volumetric flow rate at pressure Pb and temperature T,,, Mcf/D;


BI = bend index, degrees/mile; Ff = drag factor; f = Fanning friction factor; fpt =
Fanning friction factor for partially turbulent flow; fft = Fanning friction factor for
fully turbulent flow; Gg = gas specific gravity (air = 1); A t = elevation, ft (positive
for net uphill flow and negative for net downhill flow); di = internal pipe diameter,
inches; I = pipeline length, miles; p b = base pressure, psia (usually atmospheric);
pavg= average pressure, psia; p 1 = upstream pressure, psia; p 2 = downstream pres-
sure, psia; T = flowing gas temperature, OR; Tb = base temperature, OR; Z = gas
compressibility factor; p = viscosity, lb,/ft-sec; and E = effective or operating
roughness, microinches:

Steel pipe Roughness E (microinches)


New 500-700
12-months old 1500
24-months old 1750
Plastic-lined, or sand blasted 200-300

The roughness E for aluminium pipe is equal to 200 microinches.


Weymouth approximation
A very widely used formula for volumetric flow rate (Qb),
which was developed
by Weymouth, can be derived as follows:

f = 0.0035rK (8-114)

and

(8-115)
263

where M = molecular weight of the gas, lb/lb-mole; R = universal gas constant =


10.73 psia-ft3/lb-mole-"R; T= average temperature, OR; and z= average com-
pressibility factor at average pressure, pavg,and average temperature, $?
Rearranging eq. 8-115:

(8-116a)

Substituting eq. 8-114:

f = 0.0035&,

and R = 10.73 psia-ft3/lb-mole-"R in eq. 8-116a:

or

(8-116b)

Mean pressure eualuation

The mean pressure for an incompressible fluid is simp] the arithmetic average of
the inlet and outlet pressures. As discussed below, this is not the case for com-
pressible fluids, i.e., gases.
One can derive a simple formula to determine the pressure at any fractional
distance x from the inlet end in a pipe carrying gas (see Fig. 8-27).
Using eq. 8-116b, at point C where l = l ( x ) from inlet end and I = l(1 - x ) from
outlet end:

Fig.8-27. Mean pressure evaluation.


264

x +
Fig. 8-28. Pressure profile along the length of a horizontal high-pressuregas pipeline shown in Fig. 8-27.
(After Szilas, 1975, p. 36, fig. 1.2-1; courtesy of Akadtmiai Kiad6, Budapest, Hungary).

and

10.5

Equating these two equations:

p:-p,Z
-- P,Z-P22
X 1-x

Solving for p,:

p, = [ P: - x ( P: - Pz’>I0.5 (8-117)

Equation 8-117 suggests a pressure profile as shown Fig. 8-28. The mean pressure is
given by the following equation:

or

If in eq. 8-116:
265

di is expressed in inches, 1 in miles, and Qb in Mcf/D, one obtains the following


equation:

Qb [ 43.487
= (52so)y’] [
24 X 3600
1000 ] (12“))[
d8l3
w] - 0.5
(Tb’pb)
(8-119)

or

(8-120)

where Qb is in Mcf/D at Tb and p b , d , is in inches, T is in OR, p is in psia, and I


is in miles.
The following approximate forms of the Weymouth equation are also used:

(8-121)

and

(8-122)

where Q is measured in Mcf/D at the average pressure and temperature.


In eq. 8-122, the friction factor, f, has been included. It can be included in a
similar way in eq. 8-120.

The hydrate point for hydrocarbon gases (see Szilas, 1975)

Gases like methane, ethane, propane, and butane can enter the water lattice
without forming a chemical bond. Upon sufficient decrease in temperature, a solid
granular substance forms, which resembles snow or ice and is called hydrocarbon
hydrate. Besides hydrocarbon gas molecules, hydrates can also be formed by
nonhydrocarbon gases, such as nitrogen, carbon dioxide, and hydrogen sulphide.
The conditions necessary for hydrate formation and stabilization are (Szilas, 1975):
(1) The water must be in a liquid state during hydrate formation.
(2) Low temperature and high pressure.
(3) The gas must be of a covalent bond type with molecules smaller than 8 A
units.
(4) The hydrate must be water resistant.
( 5 ) The gas must be immiscible with water in the liquid state.
(6) No van der Waals’ forces should arise between the hydrate molecules.
Several methods of determining the hydrate point temperature for a given
266

pressure and gas composition have been described in literature. The Katz et al.
(1959) procedure seems to be the best suited one for natural gases devoid of
nitrogen and up to about 275 atm (4000 psi). Hydrates form when the following
condition is satisfied:

(8-123)

where y, = mole fraction of i-th component in the gas and Khi = equilibrium ratio
for the i-th component obtained from the Kh versus temperature-pressure charts
(Katz et al., 1959).
Heinze (1971) gave the following relationship for the hydrate point for nitrogen-

[
containing natural gases up to 395 atm (5800 psi):

!lyiKhi]
Th= (8-124)
0.445

where Th is the hydrate-formation temperature in OK.


Gas transmission systems
Gas transmission systems are complex network (looped or loopless) systems, flow
through which can not be treated like a simple single-line flow. Flow may be at a
high pressure with significant increase in specific volume with declining pressure, as
in the case of systems for bulk transport of gas from the field to the regional supply
stations. Flow may also be at a low pressure with negligible specific volume changes,
as in the case of supply from the regional stations to the consumers.
Although the flow in a transmission system is invariably transient, the assump-
tion of steady state flow is valid for many design and operation problems.
System of parallel lines
As in liquid flow,
(8-125)

where QA, QB, and Q , are the flow rates in lines A, B and C, respectively, and QT
is the total flow rate.
Also:
(8-126)

because the end points of all three lines are common. The total pressure drop is
given by the following equation (see Fig. 8-29):

A P T = ApA(or A P B or A PC ) + A P D
9
(8-127)
267

A
l M D
I io
!
I Section I I Section 2 ~

Fig. 8-29. Pipelines in parallel for Example 8-6.

The Weymouth equation (eq. 8-120) can be written as follows:

(8-128)

where K includes the constant terms. Thus, one can use an equivalent single line
having a length 1, and an equivalent diameter d , , that has the same capacity as the
looped line under the same pressure drop. This is given in Table 8-11.

Lines in series
A system of lines in series is shown in Fig. 8-30. In this case,
Q =Q = Q =Q
1 2 3 T

and

where Q , , Q 2 ,and Q 3 are the flow rates in lines 1, 2, and 3, respectively; A p , , A p , ,


Ap, are the respective pressure drops in these lines; A p , is the total pressure drop;
and Q , is the total flow rate.
For this system, it is very simple to define a single line having diameter d , that is
equivalent to I, miles of line of diameter d 2 and I , miles of line of diameter d 3 .

TABLE 8-11
Equivalent diameter and equivalent length for lines in parallel and lines in series

Equivalent diameter Entire line looped a Equivalent length


-
Lines
in
parallel Q
3/16
Lines in
series

a It is assumed that lengths of all lines in the looped section are equal.
268

Line I

Fig. 8-30. Pipelines in series.

Alternatively, one can determine the diameter of a line having length I,, which is
equivalent to the above mentioned lines (Table 8-11).

Example 8-6

In Fig. 8-30, assume d, = 2 in.,d , = 2.5 in., d , = 3 in.,1, = 1, = 1, = 15 d e S ,


d , = 3 in. and I, = 2 miles. Determine the equivalent ~ f : / ' / l ; / ~ratio for use in the
Weymouth equation for this system.

Solution: Refer to Table 8-11. For the looped part,


d,"/' (2)"'
-=- +-( 2 .5)8/3 +--(3)"/' - 9.446-in.8/3
(15)'12 (15)'/2 (15)'12 ft'/2

Inasmuch as length 1, = 15 miles,

d,"/' = (9.446)(15)'/2 = 36.584

Thus: d , = (36.584)3/8 = 3.857 in.


The resulting configuration is shown in Fig. 8-31. This represents two pipes in
series. Choosing an equivalent length:

( 3*8357
IDe= 2 miles x -)16/' = 7.639 miles.

Therefore, the equivalent length of the full system having diameter of 3.857 in. is
+
equal to 15 7.639 = 22.639 miles.

Alternatively:
Equivalent length of the 15-mile section as a 3-in. diameter pipe= 15 X
(3/3.587)16/' = 3.927 miles.
Hence, the equivalent length of the full system having diameter of 3 in. is equal
+
to 2 3.927 = 5.927 miles.

D C3in

t
15 miles .
1-
2 miles

Fig. 8-31. Pipeline system for Example 8-6.


269

Thus, the system in Fig. 8-30 is equivalent to:


(1) 22.64-mile long 3.857-in. diameter pipe;
or
(2) 5.93-mile long 3-in. diameter pipe.

Flow in hilly terrain


Frequently the terrain over which the pipeline is laid may not be horizontal. In
such cases, the horizontal flow equations presented for gas flow have to be modified
to account for the net head (or elevation) change from inlet to the outlet, as
described earlier for flow of liquid oil. For gases, this is complicated by the fact that
the gas properties are very sensitive to pressure.
A simple, yet valid, approach is to modify the Weymouth equation by replacing
the [( p : - p i ) / 1 ) ] term by:

PI -
1

where 1 = total length of pipeline; 3 = 2Ggz/53.33TaZa; z = total elevation dif-


ference between inlet and outlet (= zOutle, - zinlet),ft.
These relationships can be used to predict the performance of many systems.
However, it must be realized that in real situations numerical simulation techniques
may have to be employed to enable more accurate predictions. The interested reader
is referred to Szilas (1975) and other references listed for a detailed treatment of the
subject.

ACKNOWLEDGEMENTS

The authors are very thankful to Mr. Rajay K. Goyal for his valuable comments
and suggestions during the preparation of this chapter.

APPENDIX 8.1

One of the most routine calculation procedures in a gas field is the determination
of deliverability. This involves a trial and error type solution for the flow capacities
of the tubing in the well and the surface flow pipes. Iterative solutions are
introduced by the gas properties (viscosity, 2 factor, etc.), which have to be
evaluated at the average pressure. The calculation procedure is outlined below:
(1) Determine the gas flow in the reservoir:

' =
kh ( F i - P $ )
1422(pZ),,TR [1n(0.472re/rw) + S ]
(8 .I-1)
270

(2) Determine the gas flow in the tubing and casing from downhole to the
wellhead:

Q 2 = (198.6)2 (8.1-2)

Express the flow for each well.


(3) Determine the gas flow in the surface flowline from the wellhead to the
supply terminal(s):

(8 .I-3)

In all these equations, Q = gas flow rate, Mcf/D; pwr= bottomhole flowing
pressure, psia; jiR= average reservoir pressure, psia; pwh = wellhead flowing pres-
sure, psia; pd= pressure at supply end of pipeline, psia; k = reservoir permeability,
md; h = formation thickness, .ft; p = gas viscosity, cP; Z = gas compressibility
factor; Gg= gas gravity with respect to air = 1; T = temperature, OR; 1 = length of
flow conduit, ft; f = friction factor, dimensionless; S = skin factor, dimensionless;
d = diameter of flow conduit, in.; re = drainage radius for the producing well, ft;
rw = wellbore radius, ft; and s^ = 0.0375 Gglt/TaZa.
Subscripts: t = tubing from bottomhole to the surface; s = surface flowline;
av = average conditions in the reservoir; a = average conditions between the bot-
tomhole and the wellhead; and R = reservoir condition.
The usual calculation procedure is to assume that the gas flow rate, Q, is equal
throughout and establish the unknown wellhead pressure, pwh, through iterative
techniques. First, assume a reasonable value of pwh, obtain the average pressure and
temperature, and then evaluate the average p and Z for the gas. Use eqs. 8.1-1, 8.1-2
and 8.1-3 to get Q. If all three Q’s are equal, the assumption was correct. If not,
assume a new value of pwhand repeat the calculation procedure.
Dougherty (1982) presented a modified calculation procedure. He defined the
following quantities:
(1) For tubing, .the flow constant B, ( n is the number of wells) is equal to:

(8 .I-4)

and

B,, = n2B1(for n wells) (8.1-5)


271

(2) For flowline, E is equal to:

(198.6)’d:
E= (8.1-6)
G,T,ZS f s 4

(3):
kh
A, = (for 1 well) (8.1-7)
1422( P 1, TR

A, = nA, (for n wells) (8.1-8)


(4) Drainage radius is equal to:

re,l = JA77. (for 1well) (8.1-9)


and

re,n = r e J 6 (8 .I-10)
if there are n wells in the drainage area, A ft’.
(5):
K , = Xl/[ln(0.472re,l/rw) +S] (8 .I-11)
and

K, = An/[ln(0.472re,,/rw) + S ] (for n wells). (8.1-12)

(6):
C=ej (8.1-13)
Dougherty then derived his solution technique as follows:
Using eqs. 8.1-1 and 8.1-12:

P i -P$ = Q/Kn (8.1-14)


Using eqs. 8.1-2, 8.1-4, 8.1-5, and 8.1-13:

P5- CPih = (),( Q2 C-1


7) (8.1-15)

(8.1-16)
272

Adding eqs. 8.1-14 and 8.1-15, one obtains:

pi- cp&= Q/K,+ Q*(C-')/('n?) (8.1-17)

Substituting eq. 8.1-16 into eq. 8.1-17:

or

(8.1-18)

Introducing new terms a,, b,, and c:

a , = C / E + (C-l)/(?B,)

b, = lJK,

and
c = p i - Cp,2
equation 8.1-18 becomes :

a n Q 2+ b,Q - c = 0 (8.1-19)

This is a quadratic equation:

Q- [ -b,f ii6,2+]/2an

Inasmuch as negative Q is not physically possible,

\i(
I

c
Q= +)2 + a, --'an
bn (8 3.20)

Thus, Q can be easily determined. It is only necessary to assume an average


pressure for pZ product evaluation. Inasmuch as the pZ product for gases is almost
constant over small pressure ranges, any assumption of Pwh will not affect the result
significantly. If required, the calculations can be performed twice to give a better
accuracy.
This procedure is illustrated in Example 8.1-1.
273

Example 8.I-1

Calculate the deliverability of a gas field for n = 1, 10, and 20 wells if the pipeline
inlet pressure is 1250 psia. The following information is given:
A = 4000 acres; Gg= 0.75; r, = 3 in.; T, = 220°F; PR= 4500 psia; T, = 80°F;
k = 50 md; I, = 10,000 ft; h = 20 ft; I, = 5280 ft; d , = 4 in.; d , = 2.5 in.; friction
factor f = [2 log d + 6.53]-*; skin factor S = 2.0; and turbulence coefficient = 0.

Solution:
Assume pwr= 2500 psia and a common gas gradient of 0.08 psi/ft. For a depth
of 10,000 ft, the static pressure difference between wellhead and bottomhole is equal
to:

A p = (l0,000)(0.08) = 800 psi.

Due to flow, an additional pressure drop due to friction is introduced, which is


assumed to be 200 psi. Thus:
A p = 800 + 200 = 1000 psi

and
pwh=pwr- A p = 2500 - 1000 = 1500 psia.

For Gg= 0.75, ppc= 668 psia, Tpc= 406"R,

TR= 220°F = 680"R * ( TPr),= 1.675,

r, = 540"R * (Tpr),= 1.330,

-
pR= 4500 psia * ( ppr)R
= 6.737,

pwr= 2500 psia ( pPr)wf


= 3.743,

pwh = 1500 psia ( ppr),h= 2.25,


pd = 1250 psia * ( ppr)d= 1.87.

Using Fig. 8-21,

Z(TR, jj,) = 0.947,

z(TR, p w f ) =o*845,

Z ( q , pwh)= 0.690
274

and,
Z ( T , , p d ) = 0.730.

Using Fig. 8-23,


p(TR, p R ) ~ 0 . 0 2 4 8CP

and,
p ( T R , p,,) =0.020 cP.

Thus:

(0.0248)( 0.947) + (0.020) (0.845) = 0.0202 CP,


(pz)av= 2

z,= 0.845 +2 0.690 = o.7675,


T a T=+ TR
L -- 610°R,
2

s^= 0.0375Gglt -
- (0.0375) (0.75) (10,000) = 0.6007,
‘ a ‘a (0.7675) (610)

f,= (2 log d, + 6.53)-2= (2 log 2.5 + 6.53)-, = 0.0186,

f,= (2 log 4 + 6.53)-2 = 0.0167,

(198 .6),d:
B, = - (198.6),(2.5)’
GgTaZafJ, (0.75) (610)( 0.7675) (0.0186) (10,000)
= 58.9761,

B,, = (10),(58.9761) = 5897.61,

B2, = (20),(58.9761) = 23,590.43,

E = (198.6)’d: -
- (198.6),(4)’
= 1592.92,
G,T,Z,&J, (0.75)(540)(0.71)(0.0167)(5280)

A, = kh -
- (50)(20) = 0.0512,
TR (1422) (0.0202) (680)
1422(p Z)av
A,, = lOh, = 0.5120, and
h,, = 20A, = 1.024,
275

re,, = = /(4000)(43560)/~ = 7447.3 ft,

re,,, =d m = 2355.04 ft, and


re,2o= 4- = 1665.3 ft,

A1 0.0512
K, =
ln(0.472re,l/rw) +S ln(0.472 x 7447.3/0.25) + 2 = 4.43 x 10-3,
0.5120
K1o = ln(0.472 x 2355/0.25) + 2 = 0.04923,
1.024
Kzo= ln(0.472 X 1665.3/0.25) + 2 = 0.10186.
Next, the gas flow rate Q for each of the three cases can be determined as shown
below:

Parameter Number of wells


n=l n = 10 n = 20
a, = e'/E + ( eg - l ) / ( S , ) 0.0244 0.00138 0.001203
b, = 1/K, 225.734 20.313 9.817
c = p i - eJpt 17,400,953 17,400,953 17,400,953

27,102.6 112,532.45 120,338.19

- 4625.7 - 7359.78 - 4080.38


22,477 105,173 116,258

SAMPLE PROBLEMS

(1) Fracture fluid having specific gravity of 1.05 is to be injected into three wells
simultaneously as shown below. Pipe section AB is 4 in. in diameter and 500 ft long.
Sections BC, BD and BE are each 1000 ft long with I.D. of 2.5 in. Friction losses at
each wellhead are equal to 50 psi. There are 2 flanged globe valves in each one of
the sections BC, BD, and BE. The depth to the midpoint of perforations in each of
the wells is 2500 ft. The tubing is 2.5 in. I.D., and casing is 4.5-in. I.D., with
c/d = lop4. The storage is open to atmosphere (Fig. 8-32). For a fracture fluid rate
of 50 bbl/min and a pressure of 3000 psig against the sandface, determine the
horsepower of the pump required at the fracture-fluid storage outlet. Assume
viscosity of fracture fluid = 100 CP and the fracture fluid is pumped through the
tubing-casing annulus.
276

C Frocture -a
Fluid Storoge

4in 4 in ID
Pump

41:
1
Fig. 8-32. Sample problem 1.
Formation

(2) A 45"API oil with solution gas/oil ratio R,=400 scf/bbl is to be trans-
ported at a flow rate Q = 12,000 bbl/D, through a 65-mile long, 10-in. I.D. pipeline.
Pipe relative roughness, ~ / = d soil temperature T, = 50°F; a, = 6.5 X lo6
ft2/sec; inlet oil temperature = 105°F; k , = 0.5 Btu/hr-ft-OF; k , = 1.5 Btu/hr-ft-
OF; kin= 0.05 Btu/hr-ft-OF; insulation thickness = 1 in.; pipe O.D. = 5.375 in. The
pipe is buried 10 ft below the surface.
Determine:
(a) The time required for flow to reach steady state.
(b) The steady-state pipeline outlet oil temperature.
(3) Gas is to be transported through a 4.5-in. I.D. 5-mile pipeline at 60 MMscf/D.
The inlet pressure is 1500 psia, gas gravity Gg= 0.80, average flowing temperature
= 75"F, friction factor f=O.O15. For a supply pressure of 800 psia, find the
horsepower of a 90%-efficientpump that must be placed 3 miles upstream from the
exit end.

REFERENCES

Barry, E.G., 1971. Pumping non-Newtonian waxy crude oils. J. Insr. Per., 57(554) : 74-85.
Bed, C., 1946. The viscosity of air, water, natural gas, crude oil and its associated gases at oil-field
temperatures and pressures. Trans. Soc. Per. Eng. AZME, 165 :94-115.
Beggs, H.D. and Brill, J.P., 1972,An Experimental Study of Two-phase Flow in Inclined Pipes. 47th Annu.
Fall Meet., Soc. Pet. Eng. AIME, San Antonio, Texas, Oct. 8-11, SPE 4007,13pp.
Beggs, H.D. and Robinson, J.R., 1975.Estimating the viscosity of crude oil systems, JFT Forum.J . Per.
Tech., 27(9) :1140-1141.
Brill, J.P. and Beggs, H.D., 1974.Two Phase Flow in Pipes. Textbook for courses at the Univ. of Tulsa.
Brown, G.G., 1945.A series of enthalpy-entropy charts for natural gases. Trans. Soc. Pet. Eng. AIME,
160:65-76.
277

Brown, G.G., Katz, D.L., Oberfell, G.B. and Alden, R.C., 1948. Natural Gasoline and Volatile Hydro-
carbons. N.G.A.A., Gas Processors Assoc., Tulsa, Okla.
Brown, K.E. and Beggs, H.D., 1973. The Technology of Artificial Lijt Methodr, Vol. 1. Pennwell, Tulsa,
Okla., 487 pp.
Campbell, J.M., 1974. Gus Conditioning and Processing. Campbell Petroleum Series, Norman, Okla., pp.
152-185.
Carr, N.L., Kobayashi, R. and Burrows, D.B., 1954. Viscosity of hydrocarbon gases under pressure.
Trans. SOC.Pet. Eng. AIME, 201 : 264-272.
Chernikin, V.I., 1958. Pumping of Viscous and Congealing Oils. Gostoptekhizdat, Moscow, (in Russian).
Chew, J. and Connally Jr., C.A., 1959. A viscosity correlation for gas-saturated crude oils. Trans. Soc.
Pet. Eng. AIME, 216 :23-25.
Colebrook, C.F., 1938-1939. Turbulent flow in pipes, with particular reference to the transition region
below the smooth and rough pipe laws. J . Inst. Ciu. Eng. Pap. 5204, 11 : 133-156.
Craft, B.C., Holden, W.R. and Graves, Jr., E.D., 1962. Well Design, Drilling and Production. Prentice-Hall,
Englewood Cliffs, N.J., 571 pp.
Cragoe, C.S., 1929. Miscellaneous Publication No. 97. U S . Bureau of Standards.
Crane-U.S.A., 1942. Technical Paper No. 409. Crane-U.S.A., New York, N.Y.
Davenport, T.C. and Conti, V.J., 1971. Heat transfer problems encountered in the handling of waxy
crude oils in large pipelines. J . Inst. Pet., 57(555) : 147-164.
Dougherty, E.L., 1982. Aduanced Reseruoir Engineering. Lectures at the University of Southern Cali-
fornia, Los Angeles, Cal.
Edmister, W.C., 1947-1949. Hydrocarbon absorption and fractionation process design. Petrol. Eng.,
May 1947 through March 1949.
Edmister, W.C., 1961. Applied Hydrocarbon Thermodynamics, Vol. I. Gulf, Houston, Tex., 312 pp.
Frick, T.C., 1962. Petroleum Production Handbook. Vol. 11. McGraw-Hill, New York, N.Y., pp. 19-39.
Hansen, W.P., 1960. Time-saving chart for pipe wall thickness selection. Petro/Chem. Eng., 32(6) : 1-15.
Heinze, F., 1971. Hydratbildung. Lehrbogen 3/3 von der Bergakademie Freiberg.
Hydraulic Institute, 1979. Engineering Data Book. Hydraulic Inst., Cleveland, Ohio, 203 pp.
Institute of Gas Technology, 1972. Steady Flow in Gas Pipelines. Technical Report No. 10.
Kato, H., Nishiwaki, N. and Hirata, M., 1968. On the turbulent heat transfer by free convection from a
vertical plate. Int. J. Heat Mass Transfer, l l ( 7 ) : 1117-1125.
Katz, D.L., Cornell, D., Kobayashi, R., Poettmann, F.H., Vary, J.A., Elenbaas, J.R. and Weinaug, C.F.,
1959. Handbook of Natural Gar Engineering. McGraw-Hill, New York, N.Y., 802 pp.
Lee, A.L., Gonzalez, M.H. and Eakin, B.E., 1966. The viscosity of natural gases. Trans. Soc. Pet. Eng.
AIME, 237 : 997-1000.
Ludwig, E.E., 1977. Applied Process Design for Chemical and Petrochemical Plants, Vol. 1. Gulf, Houston,
Tex., 371 pp.
Makowski, M.M. and Mochlinski, K., 1956. An evaluation of two rapid methods of assessing the thermal
resistivity of soil. Proc. Inst. Electr. Eng., Oct.
Marks, A., 1978. Handbook of Pipeline Engineering Computations. PennWell, Tulsa, Okla., 347 pp.
Moody, L.F., 1944. Friction factors for pipe flow. Trans. Am. SOC.Mech. Eng., 66:671-684.
Perry, R.H. and Chilton, C.H. (Editors), 1973. Chemical Engineers' Handbook. McGraw-Hill, New York,
N.Y. 5th Ed.
Petroleum Extension Service, 1953. Oil Pipeline Construction and Maintenance. Univ. Texas, Austin, Tex.,
193 pp.
Rohsenow, W.M. and Hartnett, J.P., 1973. Handbook of Heat Transfer. McGraw-Hill, New York, N.Y.,
pp. 3-121.
Standing, M.B. and Katz, D.L., 1942. Density of natural gases. Trans. SOC. Pet. Eng. AIME,
146: 140-149.
Szilas, A.P., 1975. Production and Transport of Oil and Gas (Developments in Petroleum Science, 3)
Elsevier, Amsterdam, 630 pp.
Thomas, R. and Turner, W.C., 1953. Insulation for heat and cold. Chem. Eng., 60(6):222.
Wylie, E.B., Streeter, V.L. and Stoner, M.A., 1972. Unsteady Natural Gas Calculations in Complex Piping
'
systems. In: 47th Annu. Fall Meet., Soc. Pet. Eng. AIME, San Antonio, Tex., Oct. 8-11, SPE 4004.
Zaba, J. and Doherty, W.T., 1956. Practical Petroleum Engineers' Handbook. Gulf, Houston, Tex. 4th ed.

Potrebbero piacerti anche