Sunteți pe pagina 1din 112

UNIVERSIDAD DE ORIENTE

NCLEO DE ANZOTEGUI
ESCUELA DE INGENIERA Y CIENCIAS APLICADAS
DEPARTAMENTO DE MECNICA
FLUJO BIFSICO

HEAT TRANSFER MODEL OF FLOW IN PIPES HORIZONTAL

REALIZADO POR:

NARVAEZ ANDREA

REVISADO POR:

C.I.: 21079954

PROF.CAMARGO
LINO

PERDOMO
WILLIANNYS

C.I.: 20398097

GIPE ANDERSON

C.I.:20054894

PUERTO LA CRUZ, JULIO DE 2012

6. Heat transfer from two-phase flow


Heat transfer during the flow of gas-liquid two-phase, is found widely in industrial operations.
Occurs during the vaporization and condensation processes such as refrigerator systems, boilers,
furnaces and capacitors, as well as in applications where the heating of the two phases occurs
without significant phase shift. It has also been heat transfer two phase flow of special interest to
the nuclear energy in connection with attempts to predict the course of a loss of coolant accident.
Most of the studies published in heat transfer two phase flow has been empirical in nature and
independent flow pattern. As mentioned above, all the design variables of two phase flow strongly
depend on the existing flow pattern, including the heat transfer process. Thus, the heat transfer for
different flow patterns can differ markedly. Updated reviews of the literature on convective heat
transfer two-phase flow are given by Kinm (2000) and Manabe (2001)
6.1 Introduction
Pipe for the flow of turbulent phase, the correlations shown in Eq. 6.1 to 6.3 are commonly used to

determine the coefficient of convective heat transfer.

For the correlation of Dittus and Boelter (1930).

..(Ec 6.1)

Correlatin Colburn (1933)

...(Ec.6.2)

Correlatin Sieder y Tate (1936)

......
(Ec.6.3)

Where Nu is the Nusselt number, k is the thermal conductivity and


B W are the bulk and wall viscosities at respective temperatures,
Re is the Reynolds number and Pr is the Prandtl number, defined by

......
................................... (Ec.6.4)

Note that the difference between the three correlations is practically


negligible. A correlation was most recently issued by Petukhov
(1970)

...................... (Ec.6.5)

Where f is the friction factor. In a modified version of the correlation,


the RHS of Eq 6.5 is multiplied by the term ( / ) 0.25

Attempts to develop correlation for the two phase flow has been
carried out using equations similar to phase current. The
correlations developed are independent flow pattern. The simplest
method is the use of inhomogeneous drive model, as suggested
DeGance and Atherton (1970)

............................... (Ec.6.6)

Where all parameters are averaged on the basis of the non-sliding


retaining fluid. This approach is best suited to dispersed bubble flow
and is discussed in more detail in Section 6.3.

Davis and David (1964) suggest the following correlation is shown in


Eq 6.7 for the determination of HTP the heat transfer coefficient two
phases.

........................ (Ec.6.7)

Another approach to correlate data heat transfer two phase flow has
been the use of the Lockhart and Martinelli parameter, X Several
studies suggest a correlation follows:

................................ (Ec.6.8)

Where HL is the coefficient of heat transfer from the liquid phase


(surface) and the parameter is XTT Lockhart and Martinelli turbulent
to turbulent conditions, as shown in Eq. 6.9 (where x is the quality).

....................... (Ec.6.9)

Have proposed different values for the coefficients of Eq 6.8.


Dengler and Addoms (1956) suggest C = 3.5 and n = 0.5, while
Collier and Pulling (1962) used C = 2.5

............................ (Ec.6.10)

and n = 0.7. A similar correlation was proposed by Rounthwaite


(1968), as shown in CD. 6.10.

Notice that Eq 6.10 is equal to 6.8 Eq if neglecting the effect of


viscosity, ie, (ul / mg) 0.1 = 1, C = 1.8 and n = 0.77.
Several investigators attempted to incorporate variable two-phase
flow, such as superficial velocities or gas void fraction, the
correlations of heat transfer coefficient of two phase flow. An
example is the correlation of Dorresteijn (1970) for upstream and
downstream given vertical

by
............. (Ec.6.11)
And,

................... (Ec.6.11)

Another example of this approach is the Rezkallah and Sims (1987)


upward vertical flow correlation, given as

............ (Ec.6.12)

.................. (Eq 6.12)

Barnea and Yacoub (1983) developed a mathematical model with an


analytical solution based on the method of characteristics, to
predict the heat transfer process to unstable slug flow in vertical
upward pipe. The model provides a solution to the temperature of
the gas phase and liquid phase in function of time and the axial
position and temperature fluctuations in the pipe wall. The solutions
are presented for both constant heat flux to constant wall
temperature.
Kaminsky (1999) derive methods of calculating an average
coefficient of convective heat transfer in two-phase flow for all flow
patterns except for annular flow. These methods are coupled
explicitly with individual flow pattern hydrodynamic models for the
prediction of condensate liquid and pressure drop. The comparison
between the proposed methods: vertical and horizontal data reveals
errors of 30%. The explicit link between the hydrodynamic and
heat transfer ensures that the predictions are more robust, so that
conditions can be applied beyond the range of experimental data.

More recently there have been two experimental and theoretical


studies on convective heat transfer in two-phase flow. Kim (2000)
developed a correlation coefficients for the prediction of heat
transfer to gas-liquid flow in horizontal and vertical lines. The

correlation has been approved as published and the data collected


in this study. Manabe (2001) had acquired data convective transfer
of heat to flow from one phase and two phases. Data were acquired
in a tube of 5.25 cm diameter at 3101 kPa using crude oil and
natural gas under cooling conditions. On the basis of experimental
results, Manabe recommended Petukhov correlation (1970) for
liquid phase flow, whereas for the gas flow were suggested phase
correlations of Dittus and Boelter (1930) or Colburn (1933). Data
from two phase flow measurements include average coefficient of
heat transfer, bubble HTP vertical and intermittent streams,
horizontal and vertical annular flow and stratified flow horizontally.
The flow pattern dependent correlations of heat transfer coefficients
were presented and evaluated in relation to the data, showing
errors between 20 and 40%. As mentioned above, the following
sections present studies only dependent mechanical model or data
flow pattern of experimental local heat transfer in two-phase flow.

Stratified flow 6.2 Heat Transfer


This analysis was presented in Taitel and Dukler (1985). A heat
transfer scheme in stratified flow, including the relevant variables in
a cross-sectional area is shown in Fig 6.1. The liquid phase zone is
confined by the angle L while the gaseous phase is limited by the
angle G. higher temperatures of liquid and gaseous phases are TL
and TG, respectively. Tw () is the local pipe temperature of the wall

surface, and qw () is the local heat flux. The coefficients of local


heat transfer to the liquid phase and gas are hL () and hG ()
respectively. Note that the local variables defined, qw (), T w (),
hL () and hG () depend on the angular position around the
periphery of the tube .

6.2.1 physical phenomena in stratified flow, heat transfer gas and


liquid phases are considered separately. This is due to the fact that
the coefficient of heat transfer to the liquid phase, hL (), is often an
order of magnitude greater than the coefficient of heat transfer to
the gas phase, hG () of fact, a good approximation would be to
transfer heat to the liquid phase only because the amount of pulse
transferred to the gas phase is usually negligible. This estimate is
true for conditions of large fluid retention and may not be the case
for small water retention.

Although the heat transfer in the gas phase is negligible compared


to the liquid phase is transferred, an interesting phenomenon in
respect of the volume of gas temperatures and liquid, TG and TL.
Although only a small amount of heat is transferred to the gas
phase, there is a large increase in the temperature of the gas mass
because the heat capacity of gas is much lower than the liquid.
Furthermore, although a greater amount of heat is transferred to
the liquid phase, the temperature increase is much smaller than the
gas phase due to the heat capacity is greater in the liquid. As a
result, when heat is transferred to the two phase flow, the
temperature of the gas volume is always higher than the
temperature of the liquid volume. This applies to other types of
flows when it is not well mixed, and the region of liquid film flow of
mud. Thus, sometimes the temperature increase of the volume of
gas along the pipe can be of primary interest and importance,
therefore not be neglected in the heat transfer gas.

6.2.2 Definition of heat transfer variables. For stratified flow, as


shown in Figure 6.1 and discussed above, varying the heat transfer
coefficient around the periphery of the pipe. The coefficient of heat
transfer local to the liquid phase is defined as:

.................................... (Ec.6.13)

Similarly, the coefficient of heat transfer local to the gas phase is


defined as

..................................... (Ec.6.14)

The circumferential average coefficients of heat transfer fluid and


gas can take place in two different ways. The first method is the
average coefficient of heat transfer on the angle confined by the
area
phase, , as

.... (Ec.6.15)

The second method is based on a local average heat flow and


temperature of the wall of a phase, as shown in EQ 6.16 for the
liquid phase,

.................................... (Ec.6.16)

Where:

....................................... (Ec.6.17)

and

..................................... (Ec.6.18)

The coefficient of average heat transfer from the gas phase can be
defined similarly. In practice, there is negligible difference between
the definition of heat transfer coefficient given by Eq circumferential
average 6.15 or Eq.6.16.

Since the gas temperature is generally higher than the temperature


of liquid can occur heat transfer between the two phases. The
average heat transfer coefficient at the interface is given by

.................................... (Ec.6.19)

Where is the average heat flux at the interface.

6.2.3 General approaches. Two basic approaches may be used to


predict the heat transfer process in stratified flow. The first
approach, a simplified approach is to treat the gas and liquid phases
as a single-phase flow, using the concept of hydraulic diameter.

Therefore, the correlations available single-phase flow are such as


Dittus and Boelter (1930), which can be used to determine the heat
transfer coefficients for each of the phases. This is analogous to the
processing of the pressure drop in stratified flow, where the
calculated correlations of the shear stresses in the walls of gas and
liquid flow pipe monophasic, applying hydraulic diameters for the
phases. This approach is discussed in detail in the next section.

The second approach, a numerical approach, presented by Davis


and Guzy (1979), is similar to your solution to the pressure drop and
fluid retention. In this case, it is considered a heat flux of constant
wall, which is solved according to the temperature profile of the
liquid phase, by the expressions of Deissler and Von Karman for
eddy viscosity (see Eqs. 5.71 and 5.72). Assumed that the
temperature profile is a function only of the radial position, Roy,
normal heat flow, and that the thermal diffusivity equal to the
turbulent eddy viscosity. The gas phase, however, is treated as
single phase flow in a duct using the single-phase flow correlations.
This approach requires a numerical solution for determining the
temperature profile of the liquid phase T = T (r) and the
temperature TL of the liquid volume.

6.2.4 Simplified model. The following assumptions are made in the


simplified model:

The heat transfer coefficients are constant liquid and gas, hl =


const. and hG = const., averaged over the tube length. The
coefficients of heat transfer fluid and gas can be calculated by the
correlation of single-phase flow for the Nusselt number (as given in
Eq. (6.1 to 6.5), using the hydraulic diameters of gas and liquid
phases. Using the correlation Colburn (1933), the Nusselt numbers
(and heat transfer coefficients) for gas and liquid phases can be
determined, respectively,

...... (Ec.6.20)

and

The heat transfer coefficient of the interfaces, hi, is zero. (Another


approach is hi = hG).

Two different boundary conditions used are constant wall


temperature and constant heat flux, as detailed below.
The two different boundary conditions that can be used are shown
schematically in Fig 6.2. The case of constant wall temperature
(case a) may occur, for example, when steam condenses on the
outer tube wall. The second case (case b) is the case of constant
heat flux, which occur during the uniform heating, as by electrical
elements attached around the periphery of the pipe.

In case a constant wall temperature. Shows a diagram of the case


wall temperature constant, only the liquid phase, in Fig 6.3. As
shown in the LHS, the length of the pipe section is c and the liquid
temperature increases from the inlet temperature (known) TLI. the
outlet temperature (TBD), TLO. The temperature profile in function

of the axial length L, is shown in the RHS. The heat transferred to


the liquid phase is given by (using the log mean temperature)

................................. (Ec.6.21)

The heat transfer is related to the temperature rise by

................................ (Ec.6.22)

For a given set of flow conditions, including temperature wall, Tw


and the intake temperature, TLI, the outlet temperature, TLO QL
and the heat transfer can be determined by solving the
simultaneous 6.21 and 6.22 Eq. The heat transfer coefficient is
calculated from Eq hL 6.20 (or similar correlations). The heat
transfer gas phase and temperature can be resolved in a similar
manner.

Case B: constant heat flux. For constant heat flux is shown in Figure
6.4, only the liquid phase. In this case, as shown in the LHS, the wall
temperature and the liquid temperature increases, TWL and to Two
TLL to TLO, respectively. It is assumed that the wall temperature at
any cross section along the pipe is constant around the periphery of
the pipe. This means that the temperature of the pipe wall is a
function only of the axial position

........
................................. (Ec.6.23)

Note that this will occur for a high conductivity of the pipe wall due
to a radial heat flux to the wall. Assuming that all heat is transferred
to the liquid phase is obtained by an energy balance

........................ (Ec.6.24)

As QL, WL, SL CPL and are constant, it is assumed that the


derivative of the temperature is constant (equal LHS) and that the

wall temperature difference to liquid is constant (from equal RHS),


given respectively by,

.............. (Ec.6.25)

Therefore, Tw and TL increase linearly along the pipe, as shown


schematically in Fig RHS 6.4. integration of Eq. 6.24 is:

................. (Ec.6.26)

Eq. 6.26 to calculate the outlet temperature, TLO (from equality


LHS) and the constant temperature difference between the wall and
the liquid, Tw-TL (RHS equal). Therefore, the wall temperature of the
inlet and outlet, and Two TWI can be determined. The wall
temperature profiles and liquid are shown on the right side of Fig.
6.4.

6.3 Bubbles and dispersed flow heat transfer in bubble


The non-slip homogeneous model can predict the behavior of
hydrodynamic flow to dispersed bubble flow (see Section 2.1). In
this model, the gas and liquid phases are assumed well mixed, so
that no slippage occurs between the phases. Therefore, the two
phase mixture is treated as a single phase fluid in a pseudo speed
and the average physical properties. The physical properties are
determined from a single gas phase and the liquid properties, on
average based on liquid retention desizante not, L. the non-slip
homogeneous model can be extended to predict the heat transfer
process in the dispersed bubble flow.

This is done assuming, again, a single-phase flow and using a


pseudo-correlations of the Nusselt number for determining the
coefficient of heat transfer phase flow. Consequently, the thermal
properties are calculated from the liquid phase mixture of gas and
thermal properties, averaged on the basis of the non-sliding
retaining fluid.

The Nusselt number of the mixture can be determined from


equation modified Colbum (1933) for flow conditions of gas-liquid
non-slip, which is given by

........................... (Ec.6.6)

The thermal properties (ie, thermal conductivity and thermal


capacity of heat) are averaged based on the non-sliding retaining
liquid, given respectively by

............................. (Ec.6.27)

And

.................... (Ec.6.27)

Note that the density and viscosity of the mixture does not slip and
are defined by Eq.2.6 and 2.11, and non-slip Reynolds number is
defined by EQ. 2.16 The non-slip Prandtl number is given by

................................. (Ec.6.28)

Using Eq 6.27 and 6.28 in Equation 6.6 is possible to determine the


heat transfer coefficient mixing calculations HNS and heat transfer.
The calculations for the cases of constant wall temperature and
constant heat flux to follow a similar procedure presented above for
the case of stratified flow (see Sec 6.2.4)

6.3.1 Constant wall temperature. The heat transferred to the


mixture is given by (similar to 6.21 EC.)

................ (Ec.6.29)
The heat transfer is related to the increase in temperature (similar
to Eq. 6.22)

........................... (Ec.6.30)

The outlet temperature, and the heat transfer TMO, QM can be


determined by solving Eq. 6.29 and 6.30 simultaneously with the
heat transfer coefficient, the HNS calculated from Equation 6.6.

6.3.2 constant heat flux. For this case, the heat transfer process can
be determined by (similar to Ec.6.26)

........... (Ec.6.31)

Because QM is given for this case, Equation 6.31 allows calculation


of the outlet temperature of the mixture, TMO (equal LHS) and also
the constant temperature difference in the wall of the mixture, TWTM (from equal RHS). Thus, the inlet and outlet temperatures of the
wall, and TwO TWI also be determined.

6.3.3 Bubble Flow. The same procedure as described above, can be


used to determine the heat transfer coefficient for bubble flow. The
only difference is the use of real fluid retention can be calculated
from the bubble flow model (see Sec 4.3) instead of the no-slip
liquid retention of the above equations.

Transfer 6.4 Slug flow heat


Several studies have been published on heat transfer in slug flow
intermittently. These include Abou-know and Johnson (1952),
Akimenko et al (1970), and Zarudnev Fedotkin (1970), fried (1954),
Hughmark (1965), Johnson (1955), King (1952), Lunde (1961) and
Oliver and Wright (1964). However, the results of these studies are
presented as time and space averaged data over a slug unit.
However, the intermittent flow is inherently unstable process with
large variations with time of local flow rates, distribution of phase

and phase velocities in any cross section. As a result, therefore


expect large swings in the local heat flux at heat transfer
coefficients, and in the wall temperature. In addition, the
temperature of gas and liquid may differ substantially, which can be
important for the designer.

The unsteady process of heat transfer in slug flow is highly


dependent on the hydrodynamic flow. Dukler and Hubbard (1975)
proposed a hydrodynamic model for slug flow (see Sect 3.4.1).
Suggest that a typical slug unit consists of four zones (see Figure
6.5):
1. a swirl that is mixed in front of the slug length, Lm, where the
slow film before the slug is collected and mixed with the slug
2. the main part of the slug of liquid, Ls, where the liquid moves as
the flow occupy the entire pipe.
3. An area of liquid film, Lf, where the liquid is poured from the back
of the slug and decelerates. This fluid flows in a layered
configuration, with the height and varying speed along a distance
behind the slug.
4. A region of gas flowing over the film pocket, of the same length
LF.

The Dukler and Hubbard (1975) model (and other models of slug
flow) allows the prediction of these characteristic lengths and the
velocity distribution (ie, the slug, film speeds and gas-pocket).
Because the flow characteristics of the three zones of liquid are
different, one would expect the heat transfer process is also
different. Based on the hydrodynamic model of slug, Niu and Dukler
(1976) developed a model for

Figure 6.5. Scheme of the different areas of slug

Predict the timing and position depending on the temperature of the


fluid and wall, and the heat flux and heat transfer coefficients.
Shoham et al. (1982) measured the characteristics of local heat
transfer for flow in horizontal pipes slug. The variation in time of the
wall and fluid temperatures, heat flow and heat transfer coefficient
for the different zones were reported slug flow. Also presented a
qualitative theory to explain the substantial differences in the heat
transfer coefficients measured between the top and bottom of the
slug body. This study is presented below.

6.4.1 Experimental Program


A schematic of the experimental design is shown in Figure 6.6. It
represents a modification of the equipment, which was originally
used to study the hydrodynamics of slug flow by Dukler and

Hubbard (1975). Flow loop was 3.81 cm internal diameter and the
working fluids were air and water.

Test section. Figure 6.7 shows the thermal test section, consisting of
an inner diameter of 3.81 cm and 6.35 cm outer diameter, and a
brass tubing of 1.76 m in length, heated by electrical elements. The
tube was covered with a layer thickness of 0.5 mm asbestos and a
copper sheet 0.25 mm thick on top of which were tied heaters, this
combination of an insulator and a good conductor provided a
uniform heat flux along the pipe and around its periphery. Electric
heaters consist of tubular electric elements 27 of 2,025 W at 277 V.
They received three-phase power to 480 V thyristor controller CSCR
able to adjust the output power or 50 kW. A layer of insulation was
tied around the entire thermal section.

The measuring station was located 15 cm from the rear end of the
section of heat. In this location of cross section were measured
outer wall temperature and eight inner thermocouples, four in the
interior and the exterior wall four equally spaced along half the
periphery, as shown in Figure 6.8. Three thermocouples were
suspended in the pipe by a spider for measuring the fluid
temperature cross section of the measuring station. The inlet
temperatures of gas and liquid were measured by two

thermocouples located at the entrance of the test section. The


numbers assigned to each thermocouple is shown in Figure 6.8. All
thermocouples were 0.08 cm XacTpack type T (copper-Constantan)
being grounded junction except thermocouple 14, which had a knot
exposed to achieve a fast transient response for measuring the
outlet temperature of the gas bag.

Figure 6.6-flow diagram of experimental heat transfer (after Shoham


et al., 1982).

Fig 6.7

Figure 6.7-Schematic of thermal test section

Figure 6.8. Thermocouple locations.

The process temperature measurement is complicated by the


multichannel output variable thermocouples set in the system. A
multiplexer was constructed especially for that purpose, capable of
output channels 14 to a preselected rate up to 10,000 samples / s
sampling. The level of direct current (DC) of each thermocouple was
removed and recorded, while the level of alternating current (AC) of
each channel is amplified and then fed to the multiplexer shows all
14 channels sequentially, a second output channel provides a sync
pulse at the beginning of each cycle multiplexer. The system was
verified by the data acquisition for single phase water flow. For
these races, the outer wall temperatures IT T4 were uniform, as
were the inner wall temperatures T5 to T8. As a result, the heat flux
and heat transfer coefficient on the inner wall were constant around
the periphery of the pipe. Calculations of heat transfer coefficients
showed good agreement with the values predicted by the
correlation of Colburn (1933).

Data reduction. The 14 thermocouple outputs were sampled by the


multiplexer and recorded on analog tape. This was digitized analog
recording, and data are stored on a digital tape. Each record
consisted of a set of numbers 14, 14 corresponding to the
thermocouple readings taken during one cycle of the multiplexer.
For these tests, the sampling rate was 1,000 samples / s. Thus, the
maximum time between the two readings of the same
thermocouple was 0,014 s. This time period is small compared with
the characteristic frequency of slimy process, the time-varying
temperatures of the inner and outer walls of the tube, IT T8, were
used as the boundary condition variable in time for the solution the
equation of conduction in the pipe wall in transverse measurement,
given by

. ......................... (Ec.6.32)

where = k / Cp is the thermal diffusivity. This equation is solved


numerically using a finite difference technique assuming symmetry
about the vertical axis of the tube. Once it determines the
temperature field in the pipe, could calculate the local heat flux at
the inner surface

................................ (Ec.6.33)

Where r1 is the inner radius of the thermal section brass tube.


Because the fluid temperatures were measured at the same time, it
was determined the heat transfer coefficient local

........................... (Ec.6.33)

Where T w is the wall temperature and TB is the temperature of the


mass of gas and liquid. Were calculated heat transfer coefficients in
this way in the upper and lower positions and in two intermediate
positions around the periphery (at the location of the thermocouples
of the inner wall, see Figure 6.8). Because the axial variation in
temperature was relatively small, axial conduction was negligible.

6.4.2 Experimental Results. Thirty experimental runs were


conducted varying the mass flow rates between 0.45 and 1.6 kg / s
for water and between 0.0023 and 0.011 kg / s of air, these flow
rates cover a wide range of flow conditions slug to "blow" at which
the gas blown through the slimy liquid and the transition results in
the annular flow.

Temperature distribution. The time-varying measures the


distribution of wall temperature and a set of fluid flow conditions
shown in Fig 6.9. The energy input for this run was 8.5 W/cm2.
Temperatures are measured in time intervals of 0014 s for each
thermocouple by the multiplexer. Plotting these data produces very
smooth curves as shown in the figure, with little or no dispersion.
Shown here, for clarity, only 9 of 14 temperatures measured in a
plane normal to the axis of the pipe at the measuring station
located 15 cm from the end of the test section of heat. The data
show the following results:

1. The outer wall temperatures T1 and T4 are constant with time.


Which is expected for the wall thickness used in the pipe. Under
these conditions, changes in heat flow in the inner wall does not
reach the outer wall before initiating the next cycle of slug.

2. The temperature in the upper interior wall, T5, experience wide


fluctuations. When the gas phase passes over the measuring
station, the incoming flow in the outer wall exceeds the gas flow
transferred to the inner wall, whereby the difference in energy is
stored in the wall when a slimy liquid arrives, the wall surface is
tempered, and the stored energy is transferred to the body wall of
slime. With thin wall (low thermal capacity) of other materials of
brass (low thermal conductivity) and higher heat flux, temperature
fluctuations can be expected inner wall of over 100 C.

Figure 6.9-time variation in the measurement of fluid and wall


temperatures (see thermocouple locations Fig. 6.8).

3. Temperature fluctuations similar but lesser amplitude will be held


in the bottom of the inner wall, T8. This is because the bottom of
the inner wall is always in contact with the liquid phase. Note that
there are wide differences in temperature between the upper and
lower walls. Because the top is regularly in contact with the gas
phase, its temperature is always above the bottom wall. As a result,
one can expect to take place in the peripheral wall pipes.

4. Between slugs, the temperature of the liquid film increases as the


temperature rises bottom wall. Therefore, temperature T11
liquid film varies in phase and direction of T8, the temperature at
the bottom inner wall. Engine temperature is smaller, just after the
passage of slug and increases with time, reflecting a decrease in
heat transfer coefficient as the deceleration of the liquid film.

5. T9 shows the temperature measurement using the shielded


thermocouple located within the thermal section in a position where
it undergoes gas-liquid alternately. However, it is expected that the
shielded thermocouple has a slow response in the air. A
thermocouple junction exposed, T14 was installed at the outlet of
the test section. As shown, the two temperatures coincide with each
other after exposure times of 1 s, which is consistent with the
anticipated delay. Therefore, T14 was used as an indicator of
exhaust gas temperature real. The data show that even at these
lower heat flux, temperature of liquid and gas differ by
approximately 30 C.

6. The response time is short liquid thermocouple. Thus, the


decrease indicated T11 reflects the passage of a liquid slug of the
thermocouple. Note the small difference in temperature of the
liquid, as indicated by T9 (upper slug) and T11 (bottom of slug). This

suggests the existence of a temperature gradient in the body of the


slug.

Table 6.1. A summary of the temperature measurements that will be


useful in subsequent discussions. T5 and T8, temperatures in the
center of the top and bottom wall are the arithmetic mean of the
maximum and minimum temperatures during the passage of liquid
slugs in place. TG, was the average was obtained from the
thermocouple junction T14. TF represents the film of liquid taken
from T11 as the film of liquid which passes through the
thermocouple, where Ts is the arithmetic average of T9 and T11
taken during the entire time that the slug was present on the
thermocouples.

Heat flow. A typical example illustrates an output of a computer


showing the time variation of heat flow to the fluid in the upper and
lower inner wall (Fig. 6.10).
Unusual time intervals shown on the abscissa are related to the
frequency multiplexing, in which a unit in the X axis corresponds to

0128 s. The solid curves were superimposed on a printing machine.


For this run, the time interval between slugs was 1.92 s,
corresponding to a frequency of slug Vs = 0.52 s-1. The passage of
the slug is indicated more clearly from the heat flux at position 5,
the interior of the upper wall. It takes 0.34 s for the liquid dribble
pass through this thermocouple. At this time, the surface becomes
exposed to the gaseous phase, and heat flow drops sharply. A
pronounced peak in flow exists in the top front of the slug of liquid
due to turbulent mixing. Also shown in Figure 6.10 heat flow at the
bottom and inner wall. a comparison between the heat flow from
the bottom and which is produced at the top shows that the flow is
not uniform around the slug body. This comparison also provides a
better view of the hydrodynamic processes:

1. The effect of turbulence in the mixture in front of the slug is more


pronounced at the top of the pipe at the bottom. This is because the
upper wall surface acts as a stagnation in the mixing process.
Furthermore, the effects of turbulence mixing are not located in
front of the slug, as previously visualized, but influencing the
transfer process along the entire body of slug.
2. At the bottom of the pipe, the fluid velocity decreases
continuously and smoothly from the liquid slug to the film. The data
show that the lower heat flux also decreases continuously. The

process of change results in streams that flow around the wall and
these are often the cause of the data spread in the area of film.

Experimental Fig. 6.1O-heat flux variable in time.

Heat transfer coefficients in the slug. With a computer data output


were heat transfer coefficient at the top and bottom of the pipe,
which corresponds to the heat flow data in Fig 6.10, is presented in
Fig 6.11. This graph is typical of the results of all executions. As
expected, there is shown a heat transfer coefficient increased in the
mixing zone at the front of the slug. A nearly constant value shows
the body of the slug of liquid, while the coefficient of the liquid film
decreases as decelerates with distance behind the liquid slug, also
as expected, in regions where contact with the gas surface, the
coefficient of heat transfer decreases to a very low value. Two very
interesting results emerge from the data:

1. The coefficients of heat transfer in the mixing zone at the front of


the slug is surprisingly high. For this case, a maximum coefficient of
more than 35,000 W/m2. C was observed in the bottom of the

pipe, this value is greater than what would be expected to further


condensation of film or nucleate boiling water.
2. The coefficients in the bottom of the pipe is always higher than
the top, and in many cases, this difference can be as large as a
factor of 2.0.

Time average values of the heat transfer coefficient on the


experimentally measured area of the mixture and in the body of the
slug in the upper and lower positions were calculated and are listed
in Table 6.2.

Figure 6.11 heat transfer coefficient experimental variation in time.


The heat transfer coefficients predicted given in the table were
calculated by the correlation of Colburn (1933) (ie Eq. 6.2) applied
to the slug body, as given in the following manner

................ (Ec.6.35)

Note that Nuo is calculated for uniform wall temperature during the
passage of the slug, the physical properties of the slug is calculated
by averaging the properties of the liquid and gas phase by HLLS,
retention of fluid in the slug body gives, as shown in Eq 6.36

......... (Ec.6.36)

And

.............. (Ec.6.37)

For these calculations, HLLS was obtained from the experimental


measurements of Dukler and Hubbard (1975), the marked
difference between the coefficients of heat transfer at the top and

bottom can be easily observed in the nose (mixing zone) and the
body slug. High rates of fluid and the low gas flow rates, the ratio of
heat transfer coefficients of the top and bottom is low, as an
example, for WL = 1.59 kg / s WG = 0.0023 kg / s, this ratio is 1.5,
during these runs, the temperature variation around periphery of
the inner tube was small, approximately 6 C, when there is a
decrease in fluid flow relationship the coefficients of heat transfer
increases. In the same gas flow rate, WG = 0.0023 kg / s, but with
WL = 0.91 kg / s, the ratio is 2.3 and is approximately 3 to WL =
0.68 kg / s. Under these conditions, there is a substantial difference
in temperature between the upper and lower walls of the inner
tube, on the order of 30 C. The proportion of the heat transfer
coefficient and bottom also increases with increasing gas flow rate
at constant flow of liquid, WL = 0.91 kg / s with a low rate of gas
flow WG = 0.0023 kg / s, has a ratio of 2.3, increasing to
approximately 4 WG = 0.0091 kg / s.

The experimentally derived values of Nusselt number for slug body,


Nus = hsd / ks, in places of the top and bottom depending on the
Reynolds number of slug, Res = svsd / ms are shown in Fig points .
6.12. To calculate these parameters, the heat transfer coefficient
was obtained from the experiments, as shown in Table 6.2, the
speed was slug is Vs = VM, while the average physical properties
were calculated using Eq 6.36 , assuming a uniform value HLLS on
the top and bottom of the pipe. The number of Nussel for conditions

of uniform wall temperature, Nuo is calculated by Equation 6.35, is


shown as the dashed line. The data in the bottom of the pipe are
greater NU0, while the top are generally low.

A representation similar to the mixing zone of slug shown in Fig


6.13. For this case, the average Nusselt number, NU0, is given by
the EC. 6.35, with a coefficient of 0.023 substituting 0.03.

Fig. 6.12 Nusselt numbers for slug body.

Heat transfer coefficients in the liquid film. The heat transfer


coefficients in the liquid film varies with distance, as expected,
since the liquid is decelerated at a distance behind the slug. The
experimental results for the cases of short, moderate and long
lengths of film, Lf shown in Fig 6.14 as solid lines. In this figure, x is
the distance along the film is measured from the back of the slug.
The correlation of Colburn (1933) was used to calculate values for
comparison, using the hydraulic diameter of the film and the
properties of the liquid phase, as done in stratified flow (Eq. 6.20).
The results are shown as dashed lines in Fig 6.14. The difference in
results in the region immediately behind the slug body is given at
the bottom of the pipe, the coefficient on the back of the slug is
greater than predicted by equation 6.35, discussed above. When
areas of film are long enough, these differences disappear.

Heat transfer coefficients in the gas bag. Due to the heat flows too
low, the heat transfer coefficient in the gas bag can not be
determined with the same procedure used for the liquid phase.
Instead, we calculated the average coefficient of heat transfer of
measurements of temperature increase of the gas bag in the
thermal test section. A comparison of these measured values and
calculated from the correlation of Colburn (1933), using the
properties and the hydraulic diameter of the gas phase (Eq. 6.20), is
given in Table 6.3. Caen experimental and predicted values from 5
to 50 W/m2. C.

6.4.3. An approximate theory for heat transfer slimy flow.


Experimental data in table 6.2 and fig. 6.12 and 6.13 show that the
heat transfer coefficients significantly differ between the upper and
lower body slug. These differences may be attributed to two effects:
(1) the existence of a higher concentration of bubbles in the upper
part of the slug and (2) the existence of a significant temperature
difference between the surfaces of upper and lower wall which are
put in wetted.

Figure 6.14 Experimental and heat transfer coefficients predicted in


the region of the liquid film.

The presence of a higher concentration of bubbles in the upper


surface can be expected to decrease the heat transfer coefficient
due to its effect on the local average flow properties. However, the
heat transfer experiments was carried out in the rate of gas and
liquid, so that the slug was free from bubbles (HLLS = 1), this
showed that there are still marked differences. Moreover,
calculating the heat transfer coefficient by Eq 6.35 HLLS with
different values in the bottom and top showed that the effect of a
non-uniform concentration of gas bubbles in the coefficients of heat
transfer is small compared to the observed differences.

Fully developed flow in pipes phase, the existence of the variation in


circumferential heat flux or temperature around the periphery of
pipe has been shown to have a substantial effect on the coefficients
of local heat transfer. Solutions have been proposed for laminar and
turbulent flow in pipes (Reynolds, 1960, 1963, Sparrow and Lin,
1963; Rapier, 1972; Gartner et al., 1974; and Schmidt and Sparrow,
1978). These solutions are not applicable here because the slug
flow is mainly a problem in the region of heat input to the
temperature of the periphery of the wall. This situation exists
because it increases the temperature of upper wall during the
passage of the gas bag, just before the arrival of each slug flow.

To explain this phenomenon in a slug, is considered a short


cylindrical element of liquid placed first in the nose of the slug, the
intense mixture produces a uniform temperature, as the slime IT
moves down the tube, the cylindrical member moves back on the
nose, until it moves to the back of the slug. During this period of
residence in the slime, the element is exposed to different
temperatures around the perimeter. As a first approximation, the
wall temperature is assumed to be constant in the axial direction
(average value in the length of the slug), for any given
circumferential position. The problem is then to predict h (x, ),
where x is the distance traveled by the element in the nose of the
slug, until moving to the back, and is the peripheral coordinate.

Not yet developed a solution to this problem of pipe turbulent flow.


As a means to estimate this effect, Shoham et al., (1982) presented
a solution for the case of the laminar slug flow between parallel
plates, which are held constant but different temperatures. The
objective is to find the relationship Nu1/Nuo Nu2/Nuo and the top
and bottom plates, respectively, to explore the tendency, in
comparison with the data of the same proportions of turbulent flow
pipe. Note that NU0 is for conditions of uniform wall / isothermal top
and bottom. In this approach, the energy equation is solved for the
laminar flow of slug to a speed v between two parallel plates spaced
at a distance, e. The temperatures of the bottom and top plate TB
and TT (TT> TB), respectively, are constant in the flow direction x,

as shown schematically in Figure 6.15. The y coordinate is


measured from the bottom plate and the liquid enters at x = 0 with
a uniform temperature, TI.

Transfer Fig. 6.15 heat slug in laminar flow between parallel plates.

Heat equation for the case of two dimensions is

................................ (Ec.6.37)

where a = k / ( Cp) = / . Using an overlay technique, for


the convenience of solution (assigning 1 to the bottom plate and
top plate 2) establishing the form of dimensionless temperature
profile.

.................... (Ec.6.38)

Where

.................... (Ec.6.39)

The boundary conditions

and

........................ (Ec.6.40)

Is the dimensionless solution

....... (Ec.6.41)

where x '= (x / l) / Pe, y' = y / l and Pe = l Cp / k (Pe = l / ).


As an example, the solution of a case where the initial temperature
is 1 = 0.5 is given in Fig 6.16. Note that at x = 0.181, disappears
the derivative of the temperature in the bottom plate.

The Nusselt number in the lower and upper plates respectively, are

.................. (Ec.6.42)

Figure 6.16. Temperature profile for laminar flow between parallel


plates with different initial temperatures of 1 = 0.5.

Where = (T - TI) / (TT - IT). value of the integration can be


found from Eq 6.41 to between y 'from 0 to 1.

...... (Ec.6.43)

Because the numerators in Eq 6.42 are obtained by differentiating


Eq 6.41, the resulting Nusselt numbers depend only 1 and x ', a
typical solution for the Nusselt numbers of top and bottom,
calculated for the case where 1 = 0.5 (for the temperature profile
of Fig 6.16), this is shown in Figure 6.17. The dashed line shows the
result of the temperatures of the plate equal to Nuo, fully developed
flow between convergent plates great lengths in the value of 2 / 2,
as expected. When there is a temperature difference between the
plates, the Nusselt number of the top plate, Nu2, monotonously
decreases, as shown and finally approaches a value of 2.0 (for the
case of 1 = 0.5). However, the Nusselt number for the bottom
plate, nu1 behaves quite differently. Between the inlet region and a

value of x '= 0.09, nu1 passes through a minimum, approaching


infinity at x' = 0.09. At this point, the average temperature
approaches the temperature of the bottom wall, although the
results of energy transfer coefficient of the temperature gradient
near the wall can be expected to be larger. Thus, in most of the
region 0 <x '<0.09, nu1> Nu2 therefore the heat transfer
coefficient can be expected to be higher. For x '> 0.09, the Nusselt
number is negative, with additional heating, 1 or exceeds the
average temperature becomes higher than the lowest temperature
of the wall, although following the heat transfer fluid. For large
values of x ', the heat transfer direction changes by which the
energy moves from the bottom in the wall of the liquid, and as a
result, the heat transfer coefficient and the Nusselt number become
positive. Finally, the values of length of the plate are sufficiently
large, the two Nusselt numbers again become equal.

Figure 6.17 - Nusselt numbers at the top and bottom of slug flow
between parallel plates with different initial temperatures ( 1 =
0.5).

It is convenient to consider the results of the theory of the ratio of


Nusselt numbers, and observe Nu2/Nuo Nu1/Nuo and its variation
with respect to local temperature difference.

The results are shown in Figure 6.18. Nuo values as a function of x


'can be calculated from Eq 6.42 setting 1 = 1.0. Interestingly, in
this coordinate system, for values of x ' 0.1, the ratio approaches
a single value for a given and is essentially independent of x'.

With the premise that the area of the region behind the slug mix is
characterized by the length parameter, x ' 0.1, and O for all
locations along the slug body is now possible to compare the
coefficients of heat transfer experimental and theoretical. The
temperature data in Table 6.1 were used to calculate for each
execution, which in turn was used to determine Nu2/Nuo and nu1 /
Nuo Figure 6.18. The theoretical values resulting from Nu2 (solid
line) and nu1 (dashed line) are shown in Figure 6.12. The trend is
provided in accordance with the experiment, and especially for nu1
at the bottom, the quantitative agreement is reasonably
satisfactory. At the top, the role of gas and is especially important
for data Nu2 dispersion is more important. A similar comparison is
made in the slug flow in Fig 6.13.

Therefore, it appears that the difference in the coefficients of heat


transfer located between the bottom and the top tube observed in
the slug flow can be explained on the basis of theoretical analysis,

results from the fact that each slug is in fact a region developing
thermal input. The temperature differences which fall within the
calculation of heat transfer coefficient are substantially different
between the upper and lower walls due to the presence of a higher
wall temperature above the top of each slug. Model using a single
input stream in the region can be approximated slug correctly to the
difference in the coefficients of heat transfer in the two locations.

Figure 6.18-dependence of the ratio of Nusselt numbers of upper


and lower proportion of the local temperature difference.

6.4.4 A model of heat transfer for slug flow in horizontal pipes. Niu
and Dulder (1976) developed a model for predicting heat transfer
processes in unstable slug flow. The model allows the prediction of:
(1) the time variation in position along the pipe or (2) the axial
variation in an instant in time the temperature of liquids and gases,
temperature inside the wall at the top and bottom of the pipe and
the heat flow for each phase.

Physical phenomena. The heat transfer in slug flow is essentially a


transient process. Consider point A in Fig 6.19, which is fixed in
position on the inner wall of the pipe at any arbitrary axial location.
If a heat source located in the outer wall, then the heat transfer is

carried out in the liquid. however, the instantaneous rate of transfer


will vary depending on the fluid (gas or liquid) in contact with the
point A and in the local fluid velocities.

The shaded area bounded by a continuous curve showing the liquid


limits at time T0. At that moment, the gas passes through the
surface at point A, and the rate of heat transfer fluid is low. If the
rate of energy input to the outer wall exceeds the rate transferred
to the fluid in the inner wall, then it can be expected by increasing
the wall temperature at point A. As the slug moves downstream, the
film contains at front and emerges in the rear. Eventually, the front
of the slug will move to a position showing the limits of liquid by the
dashed line, ie at time T1. at this instant, the speed of heat transfer
(from the wall at point A liquid) is considerably greater than the
time T0 because of these reasons: (1) much higher thermal
conductivity of the liquid phase, in comparison with the gas , (2) the
surface temperature is greater under constant conditions of fluid
flow due to a previous contact with the gas phase, and (3) the liquid
temperature is lower than the gas phase. Thus, the wall
temperature falls rapidly. Thus, near the top of the pipe, the wall
temperature can be expected to oscillate with the frequency of the
slug.

Fig 6.19-Outline of a slug unit advancing on time.

At point B situated on the inner surface of the bottom wall, one can
expect a variation in the rate of heat transfer and temperature
similar to take place, although not as dramatically. At time T0, the
speed of the liquid film is low (much smaller than the slug speed),

resulting in a low rate of heat transfer, on the other hand, at time


T1 the slug of liquid is moved past the point B and heat transfer
rate is higher (approximately proportional to the square of the
speed). therefore in this location, one can expect a variation in the
rate of heat transfer and fluid temperature and the wall.
There are two additional complications associated with the unstable
hydrodynamic flow during the slimy:

1. When the gas phase passes through the point A while the liquid
film passes point B, the upper wall temperature exceeds the
temperature of the bottom wall and peripheral heat transfer takes
place through the wall of the top to the bottom of the pipe.
2. After passage of the slug, during the flow of the liquid film, the
covered portion of the perimeter of the wall by the liquid film
changes with time g. Thus, the calculation of the total energy
transferred to the gas and liquid phases requires not only the
calculated heat transfer coefficient for each phase, but, a
determination of the fraction of the perimeter covered by the liquid
at each point in space time. This information may be provided by a
hydrodynamic model drooling.

Clearly, this unstable nature of heat transfer process is


characterized not only by the behavior of the hydrodynamic slug
flow and fluid properties, but also by the thickness and thermal
properties of the pipe wall. The objective of this study is to provide
the variation in position along the pipe or the axial variation in a
time instant of the following variables:

Temperature of the liquid and gas volume.


temperature of the inner wall at the top and bottom of the pipe.
Heat flow for each phase.
This time-dependent information will allow the determination of
space and time variables or average total heat transferred to the
fluid during a specified time interval for a given set of flow
conditions, including the pipe diameter, flow, and physical
properties phase.

Focus. The methodology for the study of this simulation is shown in


Fig6.20. The input information necessary for the simulator is:

1. the data necessary to calculate the hydrodynamics of slug flow,


according to a model like the Dulder and Hubbard (1975) (see Sect
3.4.1).

Simulator Fig.6.20 flow heat transfer slimy.

2. The data needed to calculate the thermal behavior of the pipe


wall. The driving force for heat transfer is closely coupled to the
volume of the pipe wall. Therefore, it is necessary to provide the
data in the pipe wall including wall thickness, thermal conductivity
and specific heat.
3. Correlation coefficient of heat transfer between the pipe wall and
adjacent gas or liquid phases.
4. The data of the thermal boundary conditions on the outer surface
of the wall.
The simulator is built in two parts. Part A uses the model Dukler and
Hubbard (1975) to calculate the hydrodynamic characteristics of a
slug unit. Part B calculates the thermal efficiency using as input the

output of part A and the thermal input data (sections 2, 3 and 4


given above).

Can occur four different mechanisms for heat transfer in a slug unit:

1. In the mixing zone at the front of a slug, the heat transfer


coefficient to be controlled by eddies associated with the rapid
collection and mixing of the film. In this region, the swirl velocity
distributions expected for complete flow line should be significantly
distorted. The heat transfer coefficient of this zone must be much
greater, as demonstrated by the higher speed of transfer of
momentum.

2. In the body of the slug of liquid, studies show that the


hydrodynamics of the distortion in the wall is similar to that
expected in a full pipe flow. Furthermore, the gas transmission has
been demonstrated that move with the liquid slug velocity
horizontal flow, therefore, the same correlations used for heat
transfer fluid in turbulent flow pipe apply here, too.

3. In the liquid film behind the slug flow is similar to stratified


configuration. For stratified flow, the friction factor can be
calculated as an approximation, using single-phase flow correlations
in pipe using the concept of hydraulic diameter (see Sect 3.2.1). By
analogy, it is assumed that the same approach is applied to heat
transfer, this was confirmed by experimental research Shoharn et al
(1982). In their study demonstrated that, except in the region
immediately behind the slug body, the heat transfer coefficient of
the liquid film can be calculated from flow pipe correlations, such as
correlations Colbum (1933) or Sieder and Tate (1936), always used
the concept of hydraulic diameter.
4. In the gas bag on the liquid film uses the same approach as in
Article 3 (for the liquid film). Thus, one can determine the heat
transfer coefficient of the gas bag via the flow pipe correlations
using the hydraulic diameter of the gas bag.

As a first approximation, the special relationship that should apply


to the mixing zone is ignored and the form of the correlation (1933)
of Colbum for heat transfer is supposed to apply to different regions
of indications.

................... (Ec.6.45)

Where dH is the hydraulic diameter and the subscript i refers to the


location phase. Next is the application of the EC. 6.45 to the
different areas of slug:

i = l: means the slug, where DH1 is the ID of the pipe, and EC. EQ
reduced to 6.35 6.45, whereby the physical properties are
determined from the EQ 6.36.
i = 2: refers to the liquid film region behind the slug. Here, DH2 is
the hydraulic diameter of the film (which varies along the region of
film), given by the HLTBAp DH2 = 4 / SF, where SF is the wetted
perimeter of the film. For this case, EQ EQ reduced to 6.45 6.20 (for
the film) and physical properties for this case are those of the liquid
phase.
i = 3: is the gas behind the slug flow over the liquid film so that
the hydraulic diameter is DH3 = 4 (1 - HLTB) Rev / SG, where SG is
the perimeter humidified gas. EQ. 6.45 in this case, EQ is reduced to
6.20 (for gas), and uses the properties of the gas phase.

It should be noted that CD. 6.45 is based on data taken during


steady state operations, with a pipe wall temperature uniform.

Thus, application of this correlation to the slug flow conditions is


carried out as a practical approach. In reality, each slug body is a
thermal entrance region of temperature variation on the peripheral
wall of pipe. As a result, the different heat transfer coefficients
produced in the upper and lower body slug experimentally
confirmed in a later study (Shoham et al., 1982). In the region of
film of liquid / gas-hole, the conditions are also unstable. However,
the heat transfer in the region of film can be treated by an
approximation of pseudo steady state where the heat transfer
coefficient is determined by the instantaneous speed and
instantaneous average hydraulic diameter (this was confirmed by
Shoham, 1982 .)

Thermal conditions of the limit. Is necessary to specify the boundary


temperature conditions in the outer wall of the tube. In this
simulation, it is possible to configure the heat flow or the outside
temperature as constant or variable along the pipe or with time.

Modeling. The heat transfer process is modeled by an Eulerian


approach / Lagrange mixed. The pipe, whose limits are fixed over
time, is divided by a series of planes normal to the axis. This defines
M segments any of which may contain (a) a homogeneous liquid
mixture of gas if the slug is present (b) separate liquid film at the

bottom and the gas phase at the top if that particular segment is
occupied by region of the film of liquid / gas bag behind the slug.
The length of each segment, Ax, is selected so that an integer
number of segments fit into a slug unit, as shown in Fig 6.21. Within
an entire axial segment, the wall is divided into 12 sections, as
shown in Fig 6.22. The thickness of the wall is divided into three
rings of equal width and the circumference is divided into four
sectors, each including 1/4 of the total perimeter of pipe. The
advantage of the symmetry of such a division is evident.

The simulation proceeds by solving simultaneous equations of


energy to the wall and the fluid starting at t = O when a distribution
of temperature or specified heat flux is applied to the outer wall.
The initial temperature of the wall and the liquid in all segments is
set equal to the fluid inlet temperature. A computational time step,
Dt, select the At = Ax / VTB. Therefore, if the film of liquid in the
segment N (see Fig.6.21) is considered to time t, the energy balance
is calculated for a time interval Dx, where the film moves the
segment in the N + 1 time t + Dt. Thus, it is known the position of
the liquid at the beginning and end of each time interval Dt, and the
temperature change can be related to the driving forces of the fixed
locations of particular surface to which said element is exposed . in
a similar manner, the energy balance in gas phase and slug body is
executed for each control volume, and then the fluids are advanced
to the next segment.

Figure 6.21 Axial segments per unit slug breakthrough method to


control the volume.

Figure 6.22. Wall thickness segments

Energy balance for the liquid film. The increase in temperature of


the liquid film during the time Dt, N moves from segment to
segment N + 1, the result of: (1) the energy is transferred from the
inner wall, EW and (2) energy net is transferred by convection, EC.
Because the gas and liquid temperatures in the film in the region
are different, the energy transfer can also take place between the
gas phase and the liquid film through the interface. However, the
relative velocity between gas and liquid is not great, and this
transfer process is neglected. Consider the film, as shown in the
cross section of Fig 6.22, the fraction of the area occupied by liquid,

HLTB as well as all other parameters hydrodynamic hydrodynamic


model can be obtained.

1. Fluid transfer wall.


if

............. (Ec.6.46)

if

............... (Ec.6.47)

and if

(Ec.6.48)

2. Convective transfer. As a fluid element moves from segment N to


N + 1, receives the liquid moves from the element in front of him
and lose water to the element that follows, all in accordance with
the hydrodynamic model. each of these fluid streams carries energy
and the net transfer of energy contributing to the accumulation of
energy in the fluid eIemento N. The volumetric rate of liquid thrown
from anything that moves at a speed, HTLV, is (VTB - VLTB) ApHLTB.
Therefore, the net energy is caused by convection in time, Dt

(Ec.6.49)

3. Balance the equation. Transfer mechanisms contribute to


changes in temperature of the liquid film, given by EQ.6.50

............. (Ec.6.50)

Energy balance for the gas bag. Similar mechanisms for transferring
power to a gas element as it moves from the segment N to N + 1,
as presented below.

1. FLUID TRANSFER TO WALL


if

(Ec.6.51)

if

(Ec.6.52)

And if

... (Ec.6.52)

2 - convective transfer, the speed of the gas bag has been


demonstrated by: (Dukler and Hubbard, 1975)

............................ (Ec.6.53)

Where C = Co - l (Co equal to 1.2 or 2 and laminar to turbulent flow,


respectively), therefore, the flow of gas from the preceding element,
as is moved from N to N + 1, to become:

.................. (Ec.6.54)

Note that if HLLS = 1.0, this term is zero and there is no input of
convective heat transfer to the gas. In the general case,

..... (Ec.6.55)

3 - Balance Equation. The temperature change of the gas-pocket is


related to the heat transfer processes as

......... (Ec.6.56)

Energy balance for the drool.


1. Wall to fluid transfer.

(Ec.6.57)
2. Convective transfer.

..... (Ec.6.58)

3. The equilibrium equation

........................ (Ec.6.59)

Energy balance of the pipe wall. Consider a single ring in a sector of


the wall of the included angle = / 2 in segment N (eg, rings 3, 6
or 9 in Fig 6.22). Four possible transfer processes can occur: (1)
radial transfer between hoops (2) radial transfer from the inner wall
of fluids (3) axial transfer between longitudinally spaced rings, and
(4) transfer through the peripheral wall. Of course, there are
transfer from the source located in the outer wall. Consider the
outer ring segment 9, as shown in Fig 6.22. This ring segment
expands in Fig 6.23.

The energy transfer process in each time period Dt shown in EQ.


6.60 through 6.64, assuming a known heat flow in the outer wall.

.................. (Ec.6.60)

(Ec.6.61)

(Ec.6.62)
and

(Ec.6.63)

Where w is the thickness of the segment, therefore, the energy


equation for the ring element is:

For the other rings, similar equations are developed.

Calculation procedure: The calculations are initiated by setting the


fluid in all elements and the pipe wall temperature of the inlet fluid.
At time t = O, the power is turned on and started the calculation. X
= position is located so that the plane coincides with the front of a
slug at time t = o. After each time step Dt, the fluids are advanced
elements and repeat the calculations.

Fig 6.23 Energy balance of wall segment No. 9.


eventually, the fluid and wall elements reach a cyclic steady state
condition, ie, the temperature will range over time, but in any
location in space, these cycles repeten exactly. This point in the
calculation is determined by comparing the temperatures in the last

segment of a slug after two successive cycles. When this


temperature difference does not exceed a certain tolerance small ,
the computer stops and prints the results.

Results of a typical simulation. A computer program was developed


on the basis of the proposed model. As an example of the results,
consider the following conditions, which reproduce one of the
executions (1975) by Dukler and Hubbard.

WL: 2.22 lbm / s


WG: 0.00555 lbm / s

HLLs: 1
r1 (ID): 0.06 ft
vs: 1.15 s-1
Fluid: Water-Air

For purposes of these energy calculations, we selected the following


conditions.

Wall thickness: 0.01 ft


Flow in the outer wall: 10 Btu/ft2s (independent of x)
Pipe Material: Stainless Steel
Inlet temperature: 200 F (independent of t).

The hydrodynamic flow behavior was calculated by the simulator A


(see Figure 6.20) using the model (1975) by Dukler and Hubbard,
with the following results.

LV: 8.84 ft / s
LS: 2.42ft
VTB: 10.17 ft / s
Vs: 8.44 ft / s
HLTB (x): see Figure 6.24.

VLTB (x): See Fig 6.24.

To obtain these results, we selected the following increments:

X = 0,805 ft (well, 11 segments)


At = Ax / VTB = 0.079 s.

One result of this simulation are shown in Fig 6.25, where


temperatures are shown as functions of axial position. The heavy
solid curve designates the temperature of the liquid phase and the
thick broken curve represents the gas. The thin curves represent
different sectors of the wall and rings identified in Figure 6.22.
Consider first the very large differences in the temperatures of the
gas and liquid in the region of lquido-pelucla/gas-bolsillo, in this
case as much as 50 F. The fact that the temperature difference
between gas and liquid is large questions the validity of the
hypothesis that there is no heat transfer takes place between
phases.
The coupling between the wall and the fluid capacity is very clearly
seen in curve 3, which represents the temperature of the inner ring
of the upper wall. For these conditions, wall temperature oscillations

are also about 50 F and are decaying into rings that are closer to
the outer diameter of the pipe. A typical time history of
temperatures in a fixed position along the pipe shown in Fig 6.26.
The observed frequency coincide exactly with the frequency of the
slug.
Figure 6.27 shows the typical comparison between the proposed
model and experimental data reported by Shoham et al. (1982). As
shown, the pipe wall temperature are higher than expected
experimental temperatures by 25%. On the other hand, the
temperatures of gas and liquid are provided in accordance with
experimental data within 1%

Figure 6.24

Figure 6.24-values predicted by VLTS hlts and Dukler and Hubbard


Hydrodynamic mode (1975).

Figure 25.6-temperature prediction as a function of the distance


(see Fig 6.22 for wall segment locations).

6.4 Unified model for predicting flowing temperature in wells and


pipelines

A unified general equation for predicting the flow temperature in


wells and pipes, is applicable to the entire range of angles, was
presented by Alves et al. (1992). The equations are transformed
into the equation for ideal gas or incompressible liquid for wells
Ramey (1962) and the equation for pipe and Brandon Coulter
(1979) with appropriate assumptions. This study also proposes an
approximate method to calculate the flow coefficient Joule-Thomson
oil.

Fig. 6.26-Prediction of temperature as a function of time (see Fig


6.22 for wall segment locations).

Figure 6.27-Comparison of experiments and predictions of


temperature profiles (see Fig 6.8 for thermocouple locations).

6.5.1 Introduction. The temperature distribution and flow pipe wells


often predicted by different methods. The method normally used

Ramey (1962) to predict the temperature distribution (vertical) of


the well. This method incorporates rigorously the complex process
of transient heat transfer between the well and reservoir. Method
Ramey, however, is limited to ideal gas or incompressible fluid. The
equation of Coulter and Bardon (1979) is commonly used for the
prediction of the temperature (horizontal) of tubing. It takes into
account a more rigorous thermodynamic behavior of the fluid,
entering the Joule-Thomson coefficient. Although equation of Coulter
and Bardon (1979) was derived originally for the gas flow can also
be used for the flow of liquid phase or two phase flow.

The use of different methods for pipes and shafts, each of which has
its own limitations, it is undesirable for the practical design. This
chapter presents a general and unified equation for predicting the
temperature of the flow. Can be applied to pipes or injection wells
and production under single or two phase flow, especially angle
between models from horizontal to vertical, with models of fluid or
oil composition.

Previous studies 6.5.2. Predicting a rigorous temperature


distribution flowing wells and pipelines is complex. Requires the
simultaneous solution of mass, momentum and energy conservation
equations. The solution is complicated by the thermal interaction

between the flow and the environment, especially in the reservoir.


Thus, a rigorous analysis for the solution of this problem is not
currently possible. In the past, attempts have been numerical
algorithms or approximate analytical solutions.

The numerical algorithms apply a double iterative procedure,


temperature and pressure to solve the three conservation equations
simultaneously, which requires knowledge of the thermodynamic
behavior of the fluid. Gould (1979), Gregory and Aziz (1978),
Furukawa et al (1986) and Goyon et al (1988) proposed these
computational algorithms.

Several researchers have proposed solutions for predicting


approximate analytical temperature. Explicit expressions for the
temperature distribution of the fluid is obtained by making strong
assumptions about the geometry of the pipe, heat transfer with the
environment and the thermodynamic behavior of the fluid.

Schorre (1954) conducted a pioneering study for the prediction of


temperature on horizontal gas pipes. Its equation for explicit
temperature results in a continuously decreasing temperature
profile along the pipe, it never reaches an equilibrium value. Coulter

and Bardon (1979) modified the equation Schorre. The modified


equation predicts a temperature profile of fluid which asymptotically
approaches a temperature slightly below the surrounding
temperature. In the equation of Coulter and Bardon (1979), the
thermodynamic behavior of the fluid is taken into account more
rigorously. However, the assumptions of the heat transfer in steady
state with a constant temperature environment and limit horizontal
flow with this method are made to a single pipe.

Ramey (1962) proposed a new method for predicting the


temperature in wells. This method coupled with the mechanisms of
heat transfer in the well with the transient thermal behavior of the
reservoir. The equations of temperature for the injection of hot
incompressible liquid phase or single phase ideal gas flow have
been derived. Satter (1965) later included the effect of changes
phase during operation of steam injection. In Ramey method, the
transient thermal behavior of the reservoir is determined by solving
the problem of radial heat conduction in an infinite cylinder.
Resistance to heat flow in the well, caused by the presence of the
tube wall, pipe insulation, fluid in the crown shell and tube wall and
the casing and cement, are incorporated in a heat transfer
coefficient in general. Willhite (1967) proposed a method for
determining this coefficient. Shiu and Beggs (1980) developed an
empirical correlation for the production of 10 oil wells to determine
the distance of relaxation defined by Ramey.

Their work is really an attempt to avoid the complex calculation of


the coefficient of total heat transfer in the well and the behavior of
the transient heat transfer reservoir. Although this correlation
simplifies the method of Ramey, caution should be used as an
approximation. Sharma et al (1989) modified the Ramey equation
for the case of producing wells drilling with a heater. Finally, Sagar
et al (1991) developed a simplified method for predicting the
temperature suitable for hand calculations on the basis of field data.
All these methods include strong assumptions relating to the
thermodynamic behavior of fluids and, therefore, are applicable
only to limited operational conditions.
6.5.3. Model development. The derivation of the general and unified
equation for predicting the temperature is performed by applying
the conservation laws of mass, momentum and energy balances on
a differential control volume of a pipe. The resulting differential
equation is integrated under simplifications, however sounds like a
guess. Applying the steady-state mass, momentum and energy
balances in the differential control volume leads to

(Ec.6.65)

(Ec.6.66)

and

(Ec.6.67)

Where Q is the heat flow and e is the internal energy per unit mass.
Using the mass balance (Eq. 6.65) can be reduced further in Eq 6.66
and 6.67, respectively,

(Ec.6.68)

(Ec.6.69)
0

(Ec.6.70)

The transfer of heat to the surroundings can be expressed by the


concept of heat transfer coefficient overall U as

Equation 6.71

Where TE is the temperature of the external environment. Must be


carefully considered the geometrical configuration of the widenings
or pipe and all heat transfer mechanisms involved in the
determination of U. It is a general expression for U

Equation 6.72

For the prediction of the temperature widenings, the first term


within the brackets, 1/Uo represents the heat transfer mechanisms
in the well, while the second term, f (t) / kE represents the transient
heat transfer in the reservoir, for predicting pipe temperature is
generally considered the first term only. Equation 6.70 and 6.71 can
be combined to produce

Equation 6.73

At this point, the analysis is rigorous and similar to that presented


by other authors. Temperature prediction can be carried out strictly,
if so provided a method for determining the enthalpy of liquid in
each condition of pressure and temperature. This is the case, for
example, when known fluid composition and may be generated by a
simulator thermodynamic Vapor-liquid-equilibrium (VLE) an enthalpy
chart. In many cases, this information is not available, and
recommended another approach. The enthalpy gradient can be
written in terms of temperature and pressure gradients as

Equation 6.74

6.73 and 6.74 Eq combining results in

Equation 6.75

6.75 Eq can be rewritten as follows most convenient

Equation 6.76

Define a relaxation distance, A, proposed by Ramey (1962).

Equation 6.77

and a dimensionless parameter, as

Equation 6.78

Equation 6.76 can be rewritten as

Equation 6.79

So far, only math maneuvers have been made to the enthalpy


equation, and the analysis was done rigorously without
simplification. Now, assume that the surrounding temperature is a
linear function of depth, which can be expressed as

Equation 6.80

Substituting Equation 6.80 into Equation 6.79 gives

Equation 6.81

If, for a given tube segment, U, Cp, , gE, 0, vdv / dL, and dp / dL
can be considered approximately constant, Eq 6.81 can be
integrated, yielding an explicit equation for temperature, originally
developed by Alves (1987).

Equation 6.82

Equation 6.82 is general and can be applied to any angle for singlephase flow or two phases. The average values of Cp and N, and the
pressure derivative dp / dl; depends on the average pressure and
temperature of the tube. Thus, an iterative procedure is necessary
for the calculations, the pressure gradient can be determined by
any method of two-phase flow. For fluid compounds may be
generated tables to give enthalpy values of Cp and . To the crude,
however, a method is necessary to approximate the values of these
parameters.

6.5.4 Approximation for oil. For the more general case of a twophase flow, the total enthalpy of a mixture is the sum of the
enthalpies of the individual phases. Therefore, the derivative can be
written as enthalpy

(Ec.6.83)

..................... (Ec.6.84)

The enthalpy derived for each stage is given by

Equation 6.85

and

Equation 6.86

Where

Equation 6.87

And
(Ec.6.88)

Are obtained by applying the thermodynamic behavior of real gas


and liquid as incompressible fluid assumption

Equation 6.89

and

Equation 6.90

Therefore, Equation 6.85 and 6.86 become

Equation 6.91

And

Equation 6.92

Substituting Eq 6.91 and 6.92 in 6.84 yields equalization

Equation 6.93

Rearranging Equation 6.93 gives

Equation 6.94

So for a mixture of two phases, as expected, the mean heat


capacity is

Equation 6.95

and becomes the Joule-Thomson coefficient average

Equation 6.96

Equation 6.97

To a liquid phase NS L = 1 and = L> 6.97 Eq is the same as


Equation 6.90 for an incompressible fluid. Also, for a gas phase, L
= O and Eq NS g = 6.97 is equivalent to Equation 6.89 for a real
gas. The heat capacity of water and hydrocarbons do not vary
greatly over a wide range of temperatures. Therefore, average
values of Cp can be easily obtained and used for any widening or
pipe. Correlations for the gas compressibility factor, z, are available
in the literature. Therefore, the values of n can be easily evaluated
with 6.97 EQ. As can be seen, very good approximation of Cp and 1)
can be determined without any assumptions severe.

6.5.5 Results and discussion. An important feature of the overall EQ


6.82 is that degenerates into the equation (1979) Coulter and
Bardon (pipe) or equation Ramey (1962) (for ideal gas or

S-ar putea să vă placă și