You are on page 1of 69

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/312044919

MSc. thesis: The Effect of Surfactant on Flooding Phenomena in Vertical,


Counter-Current, Gas-Liquid Flows

Thesis · November 2016

CITATIONS READS

0 612

1 author:

Sourojeet Chakraborty
University of Toronto
5 PUBLICATIONS   20 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Understanding the phenomena of wetting and coalescence in complex fluids View project

All content following this page was uploaded by Sourojeet Chakraborty on 04 January 2017.

The user has requested enhancement of the downloaded file.


Department of Chemical Engineering
Imperial College London

Sourojeet Chakraborty
CID: 01088178

MSc. Research Project


The Effect of Surfactant on Flooding Phenomena in Vertical, Counter-Current, Gas-Liquid
Flows (part of a SHELL UK funded project)

Primary Supervisor
Prof. Omar K. Matar
Professor of Fluid Mechanics
Imperial College London

Co-Supervisor

Dr. Christos N. Markides


Reader in Clean Energy Processes
Imperial College London

A Master’s thesis submitted to Imperial College London to fulfil the requirements for a MSc.
and DIC Degree in Advanced Chemical Engineering of the Imperial College London.

Declaration of Originality

I hereby certify that the material in this submitted Master’s thesis is my own work. Any work
that is not mine, has been properly acknowledged and referenced.
2

1 Abstract*

Qualitative behaviour and effects of surfactant are investigated on the interfacial dynamics, wave
characteristics, formation and dissociation of bubbles/droplets, regime transitions and flooding in
vertical, counter-current, gas-liquid flows through exhaustive experimental campaigns. High speed
axial imaging is employed to track regime behaviour, with Exxsol- D80 (a light oil) and Baker Petrolite
(a non-foaming surfactant); air being maintained as the gas phase. Surfactant free runs are performed
for benchmarking against surfactant laden runs taken at varied concentrations, on a 1.44 m long, 40
mm bore Perspex tubing. Imaging of the lower test section leads to clear identification of different flow
regimes viz. downwards, flooding, bridging/ break up, standing waves and dry out. The surfactant
shifts regime boundaries by smoothening the flow but cracks the test section over prolonged timescales.
Candidate mechanisms proposed to explain the interfacial topologies include tip-streaming, interfacial
rigidification and stress reduction due to Gibbs-Marangoni effect induced surface tension decrement.
Preliminary quantitative investigations of the flooding point and the liquid film thickness are
performed; the latter being statistically validated. Experimental data extracted for the air-oil system fits
semi-empirical correlations for air-water systems satisfactorily; these two systems are compared,
contrasted and results discussed, with future suggestions.

*Part of a conference paper titled ‘Flow behaviour and transitions in surfactant-laden gas-liquid vertical flows’, to be presented
at the 69th Annual Meeting of the APS Division of Fluid Mechanics at Oregon, USA
3

Contents

1 Abstract……………………………………………………………………………………………….2

2 Introduction and Context…………………………………………………………………...………..9

3 Literature Review………………………………………………………..….………………………12

3.1 Flooding: Previous Research Insights…………..…….……………………………………….………......12

3.2 Suggested Flooding Mechanisms…………..…….……………………………………..………..….........13

3.3 Parameters Affecting the Flooding Onset……..…….……………..…………………..………..…..…....15

3.4 Mathematical modelling of the Flooding and Post-Flooding Onset……….………..………....…...…..19

3.5 Mathematical characterisation of Flow Regimes in Flooding………………………..……….…...…....23

3.6 Surfactants………………………..…………..…….……………………………………..………...............27

3.7 Foaming in Surfactants……………………..…….……………………………………..………..…....…..29

3.8 Effect of Surfactants on the Flow Morphology and Pressure Drop….……………..……….…..............30

3.9 Further Research Scope: Possible Areas of Study..…….………………………..……………..…...…....35

4 Project Objectives…………………………………………………………………………………...36

5 Experimental Methodology…………………………………………………..………………….37

5.1 Rig Revamping…………………………………..…….………….……………………..….……..………..37

5.2 Materials………………………….…………..…….…………….……………………………….……..…..38

5.3 Rotameter Calibration…………………………..…….……………...…………………..……..…...……..39

5.4 Experimental Runs……………………………………………………..……….………..………......……..40

5.5 Experimental Setup…………………………………………….……….……………...…..……….…..…..41

6 Results and Discussion…………………………………………………….….…………………42

6.1 Fitting of Experimental Data………………………..…….………….…….……………….……………...42

6.2 Flooding Profile: No Surfactant………………….…………….………………………..……….………...43

6.3 Flooding profiles for different surfactant concentrations……..…….………………..……….………...43


4

6.4 Interfacial Topology and Regime Transitions…………..…………………….…….…..………………..44

6.5 Flow Smoothening, Drop and Bubble Reduction at Flooding……………………..………….…….…..46

6.6 A Preliminary estimate to the Liquid Film Profile…………………………………...…………….…….47

6.7 Discussions……………………………………………………….…………………….……..…….…..…..49

7 Conclusions……………………………………………………....…………………………………50

8 Acknowledgement………………………………………………...………………………………..52

9 References…………………………………………………………………………...……….……………….53
10 Appendix……………………………………………………………………………...……………60

10.1 Program in Dev C++ Ver 5.4.2 for data abstraction……………….....…………………….……...…..60

10.2 Sample Output for the program presented in 10.1………………………....………………….…........63


5

List of Tables and Figures

Tables

Table 1: Characterization of Flow behaviour in air-water multiphase systems, subject to surfactant


addition………………………………………………..…….…………………………………………………....29

Table 2: Comparative study of the various set-ups established by various researchers to understand flow
behaviour in the presence of surfactants………………………………….....……….……………......................31

Figures

Fig 1: Schematic representation of the formation of a Taylor bubble prior to flooding (reproduced from Jayanti et
al., 1993)………………………………………………………………………………………………...………14

Fig 2: Qualitative profiles depicting the flooding behaviour with change in the vapour and liquid flow rates.
[Reproduced from http://thermopedia.com based on Mc.Quillan and Whalley (1984)]………………...….22

Fig 3: Comparison of flooding data for sharp-edged, bell-shaped (taped) and porous wall liquid exit [reproduced
from Govan et al. (1991)]……………………………………………………………………………………..….23

Fig 4: Flow regimes without and with Trifoam 800-Block surfactant, at 1000 wppm. [reproduced from van
Nimwegen et al. (2015)]……………………………………………………………………………………...…..33

Fig 5: Influence of surfactants on the Air-Water System flow and Morphology [reproduced from van Nimwegen
et al.,(2015)]……………………………………………………...…………………………………………...….33

Fig 6: Gantt chart showing the expected work plan leading to completion of the MSc. thesis……….........………36

Fig 7: Reduction in the surface tension with increase in the defoamer concentration of Baker
Petrolite………………………………………………………………………………………………………..…38

Fig 8: Temporal variations in the surface tension values for Baker Petrolite studied at different surfactant
concentrations………………………………………………………………………………………………39

Figs. 9(a),9(b): Calibration curves for the rotameters used in the Upflow Rig setup with respect to water….…39

Fig 10(a), 10(b): Comparison between the calibration curves for the rotameters used in the Upflow Rig setup
with respect to water and Exxsol D80………………………………………………………………...…………40

Fig 11: Schematic of the experimental setup used for experimental runs………………………..……....….....41

Fig 12: Comparison of experimental results of the superficial liquid velocity with correlations proposed by
Wallis (R2=1) and Sun-Kutateladze (R2=0.885) as statistically estimated from SigmaPlot 13.0……...………..42

Fig 13: Flooding onset profiles without any surfactant loading, compared to a two-period, moving average time-
series model (best estimate) ……………………………………………………………………………………...43

Fig 14: Flooding profiles as a function of the non-dimensional, superficial velocities, plotted at different surfactant
loadings…………………………………………………………………………………………………………..44
6

Fig 15: Qualitative high speed images of the interfacial topology under different experimentally tested flow
conditions……………………………………………………………………………………………….…..……45

Fig 16: Qualitative effects of surfactant addition on the air-oil multiphase system at the flooding point…....….46

Fig 17: Variation of the liquid film thickness with the liquid phase non-dimensional superficial velocity for the
“no-surfactant” case……………………………………………………………………………………….……..47

Fig 18: Statistical fitting of the film-thickness data on SigmaPlot 13.0……………………….......…………….48

Fig 19: Corrosive action of surfactant Baker Petrolite on the Perspex tube of the rig…………………………...49
7

Nomenclature: List of Symbols and their meaning

Symbol Nomenclature Symbol Nomenclature


∆p Pressure drop μ Absolute viscosity
L Flow length Dp Particle diameter
Void fraction ρ Fluid Density
vs Superficial velocity δRMS RMS Film thickness
δ(t) Film thickness <δ> Time averaged film thickness

g Acceleration due to gravity 𝑈𝐺∗ Dimensionless superficial


gas velocity

𝑈𝐿∗ Dimensionless superficial liquid 𝜕 Partial derivative symbol


velocity
sin() Natural sine function 𝑈𝐺 Superficial gas velocity
𝑈𝐿 Superficial liquid velocity L Liquid flow rate
G Gas flow rate D Flow diameter
Fr Froude Number Bo Bond Number
σ Surface tension 𝜌𝐿 Liquid phase density
𝜌𝐺 Gas phase density 𝑄𝐿 Same as L, the liquid flow rate
tanh () Hyperbolic tangent function HLB Hydrophile-lipophile balance
MW Molecular mass of the whole MWh Molecular mass of the hydrophilic
molecule portion of the molecule
m Number of hydrophilic groups n Number of lipophilic groups
Hh Value of the hydrophilic groups Hl Value of the lipophilic groups
𝑢𝑐 Gas velocity at the onset of loading 𝐶𝐷 Drag coefficient
𝑊𝑒 Weber number ∆ 𝑃𝑓 Frictional pressure drop per unit
𝐿 length
𝑢𝐴 ∆𝑃 Type A standard uncertainty 𝑢𝐵 ∆𝑃 Type B standard uncertainty
from pressure drop measurements from pressure drop measurements
𝑢𝐴 (𝐻𝐿 ) Type A standard uncertainty from 𝑢𝐵 (𝐻𝐿 ) Type B standard uncertainty from
liquid holdup measurements liquid holdup measurements
k Coverage factor for uncertainty U Expanded uncertainty
𝐷𝑒 Equivalent diameter 𝑟32 Sauter radius
PI Polydispersity Index 〈𝑟〉 Mean radius of bubble
V* Flooding velocity τw Wall shear stress

τi Interfacial shear stress ANOVA Analysis of Variance


𝜂𝐿 Kinematic viscosity 𝐺𝑟𝐿 Effective liquid Grashof number
8

𝑓𝑖 Interfacial Darcy-Weisbach friction 𝜋 Multiplicative functional relation


factor
𝑢+ Velocity parameter 𝑦+ Distance parameter
𝑢̈ ∗ Friction velocity α Post flooding void fraction
CMC Critical Micelle Concentration 𝜇𝐿 Absolute viscosity of liquid
𝑂ℎ𝐿 Liquid phase Ohnesorge number z Flow or propagation axis

𝐷∗ Non-dimensional tube diameter 𝛿𝑡ℎ Non-dimensional theoretical film
thickness
𝑅𝑒𝐺 Gas phase Reynolds number 𝑑𝑃 Pressure gradient along flow axis

𝑑𝑧
𝐾𝑢𝐺 Gas phase Kutateladze number 𝐾𝑢𝐿 Liquid phase Kutateladze number
𝜃 Inclination angle of tube H Height of tube
𝛿𝑝𝑓 Plug flow liquid film thickness 𝑝(𝑣) Initial wave distribution function
in the annular flow regime
𝑒𝑟𝑓 Gaussian error function 𝜆 Arithmetic Mean
𝑠𝑑 Standard deviation r Pipe radius
𝐹𝑟𝐿 Liquid phase Froude number 𝑙𝑠 Minimum slug flow length

𝑛 Non-dimensional gas phase initial βci Wave growth rate


flow velocity
RMS Root Mean Square CDF Cumulative Distribution Function
9

2 Introduction and Context

Vertical, counter-current gas liquid flows are frequently encountered in all major chemical processes
involving multiphase flow. If gas flows upwards in a vertical tube and if liquid is introduced from the
top into the tube, it is evident that the type of liquid flow observed will depend on the gas velocity. For
zero or particularly low gas velocity, the liquid will fall down the tube in the form of a thin film; and
theoretically speaking, if the tube ends are stretched at both ends infinitely, then this flow would
essentially continue indefinitely. This is however the ideal case, in practice what happens is that as the
liquid reaches the bottom end, the gas tends to disrupt the uniformity in flow and pushes the liquid
upwards. Thus for real tubes, there is always a competition that exists between the liquid and gas flow
rates at the gas inlet section. With increase in the gas flow rate, the effect becomes more pronounced. If
the gas velocity is kept on increasing for such a system, a point will eventually reach when large
bridging waves developed in the test section, essentially carries the liquid film up above the injection
point. This point is commonly referred to as the ‘flooding point’, and is traditionally seen as a point
that represents the onset of competition between two liquid films- one falling and one climbing.

Consider that the gas velocity is now increased to a magnitude sufficient enough to essentially push
the entire liquid upwards via the mechanism of a climbing film. Now, if the gas velocity is reduced, a
unique flow condition will be exhibited wherein the liquid phase, in addition to propagating upwards,
also begins to creep down the test section walls from the liquid injection point- this point is known as
the ‘flow reversal’ point. One can hence see that such alteration in the gas velocity, in the manner
mentioned before, will lead to identification of these two unique flow points. Past research has
conclusively demonstrated that while boundaries determined by the above procedure are reasonably
sharp, a considerable hysteresis exists between the two points, and that these two behaviour conditions-
one with climbing film and the other with falling film, are not to be thought of as occurring at the same
flow conditions. Thus, at the flooding onset, the column which was a gas-continuous, liquid-dispersed
equipment, changes to behave as a liquid-continuous, gas-dispersed equipment. Flooding essentially
leads to a flow behaviour reversal.

Since the phenomena’s discovery in the last century, researchers have attempted to understand
flooding behaviour, both experimentally and computationally. One of the major grounds for criticism
in this area of research is the experimental design. The earliest experiments (Wallis, 1961) simply
employed a tube leading into tanks at both ends. Water is fed to the top tank and air to the bottom.
However, it was conclusively proved that such a simplified experimental system as used by Wallis,
would give rise to other undesirable, extraneous effects such as inlet instabilities, owing to end effects.
Thus, for the purpose of understanding the behaviour, it is desirable to design a system that creates a
smoothly injected liquid film. Most noteworthy of these attempts have been the usage of overflow weirs
10

to introduce the liquid (Nicklin and Davidson, 1962) and the use of a porous sinter (Hewitt and Wallis,
1963); the latter is definitely more likely to control the liquid flow, without causing the problem of
“lipping” in weirs. Yet another problem is the successful separation of liquid and gas phases just prior
to the gas entry point: Nicklin and his co-researchers devised a bell contoured outlet for the gas velocity
to rise uniformly, prior to its interaction with the liquid phase. The entrance section was made wooden
to enhance wetting. While their attempt is certainly noteworthy, this may in turn lead to the formation
of spurious standing waves at the bottom of the test section, which is most certain to cause premature
flooding. Hewitt and Wallis devise a more satisfactory solution to tackle the problem of entrance. In
their experimental setup, the liquid film was sucked off after a given length through a further porous
wall section, thus enabling stabilisation of the gas velocity profile in the dry section immediately
following the water exit. However, despite the Hewitt and Wallis experimental setup being generally
the most satisfactory one, the setup demonstrates some limitations-most notably the range of
parameters investigated have been critiqued as too less and somewhat limited. For instance, the
distance between the sinter sections of the inlet and outlet has a considerable impact on the flooding
rate- something which Nicklin and Davidson erroneously concluded was due to effects arising from
the relative roughness of the test section material.

Theoretical attempts to correlate and predict regime transitions as described above have, in general,
not been very successful. Early efforts (Davidson and Shearer, 1965) include balances on the pressure
resulting from liquid phase gravity and curvature effects and surface curvature effects within the liquid
phase, caused by changes in the gas flow direction. Alternatively, a generalised Navier-Stokes
equations have been employed (Schutt, 1959) to furnish a solution suggesting that transitions to
flooding will occur primarily due to effects of interfacial shear. Schutt’s results, while extremely
difficult to apply, agree qualitatively with observed results. The empirical correlation of Wallis led to
further work (Sun, 1979) that incorporates the non-dimensional Kutateladze numbers. The Wallis and
the Sun-Kutateladze correlations have been most extensively used to fit flooding data. A fairly
consistent condition for flooding (Pushkina and Sorokin, 1969) is when the gas phase Kutateladze
number is 3.2; this value has been obtained through an exhaustive database of experimental runs.
Attempts have been made in more recent times (Cioncolini and Thome, 2010) to predict prior-flooding
entrainment by 9 empirical correlations, finding that it is the core liquid Weber number that
dominantly predicts the entrained fraction. Loading has been extensively studied (Waltrich et al., 2015)
in such equipment. The model presented by these researchers presents simulation results for the
axial liquid distribution for transient gas-liquid flows in such channels. Transition from annular-
to-churn flow is seen to exhibit pseudo-steady state flow.

Recent experiments have been conducted by van Nimwegen et al. (2015) investigating the qualitative
behaviour of surfactants on the air-water multiphase system using high speed imaging techniques. In
11

a separate paper, the researchers investigate the morphology of air-water vertical counter-current
systems with different surfactants and conclude that despite using different surfactants, the qualitative
behaviour is similar. Their research conclusively demonstrates that surface active agents would reduce
the surface tension for the system, thus altering the regime boundaries; however it must be stressed
that a mechanistic model for the air-water system with or without surfactants has till date, not been
established. It is thus, the need of the hour to examine such multiphase systems, and to properly
understand the role of surfactants in flooding and regime transitions in such vertical, counter-current
multiphase systems.

In this present work undertaken at Imperial College London, qualitative investigations have been
undertaken for the oil-air system, with and without surfactants, at various concentrations. The present
research attempts to understand not only the qualitative behaviour of regimes in surfactant free and
surfactant laden multiphase systems, but also attempts to understand their effects on the flooding
characteristics. Distinct transitions are seen to happen upon surfactant addition into the oil, and for the
same liquid superficial velocity, transition to a new regime happens at a lower gas velocity. These
effects would be described more comprehensively later. High-speed imaging of the test section
conclusively demonstrates that the interfacial topology is in fact, significantly altered by the addition
of surfactant into the oil-air multiphase system. The empirical correlation of Wallis is employed to get
a preliminary understanding of the liquid film thickness with increase in the non-dimensional
superficial liquid velocity. These results are subject to future validation by more quantitative,
mathematically rigorous tools; nevertheless a qualitative understanding of the changes occurring
across flow regimes is the first step to understanding and comprehensively describing the phenomena.
12

3 Literature Review

3.1 Flooding: Previous Research Insights

Sherwood et al. (1938) first investigated flooding and proposed that the phenomena could be correlated
to superficial gas velocities in both random packed and well stacked columns. Uchida and Fujita's
(1939) unpublished results on flooding velocities post exhaustive experiments led Sherwood to state
that dumped ring correlations may be satisfactorily employed to obtain flooding velocity profiles for
previously undocumented packing. Kamei et al. (1954) performed experiments in tube diameters
ranging from 19 to 49 mm in a 2.5m length test section to present a non-dimensional correlation
predicting the flooding onset. Cetinbudaklar and Jameson (1969) investigated flooding mechanisms in
air-water based vertical counter-current two phase systems and deduced the formation of waves,
progressively increasing with time, as an indicator of the flooding onset. They deduced that when the
wave growth rate βci >0, the wave growth curve exhibited a velocity V* called the flooding velocity.
Their predicted flooding velocities are well in accordance with the results of Clift et al. (1966) and Cohen
and Hanratty (1968).

Theoretical models of gas-liquid flow was first conclusively studied by Wallis and Makkenchery (1974)
where they analysed two-phase flow using techniques of averaging, continuum models and ad-hoc
based models. Suzuki and Ueda (1977) highlighted experimental results on flooding correlating the
flooding parameters of liquid flow rate, liquid viscosity, tube diameter, tube length and surface tension.
Fairly good empirical correlations are presented for various tube lengths for flooding, employing the
previous experimental data of Kamei et al. (1954), Kusuda (1974), Levich (1962) and Portalski (1972).

Wallis analysed the phenomenon of flooding onset and reported that under certain conditions, a liquid
film with a zero net flow rate may be held stationary by the upwards flowing gas. He explained this
"critical" gas velocity in non-dimensional terms, however Pushkina and Sorokin (1969) suggested a
much better criteria using the Kutateladze number.

1
𝐾𝑢𝐺 = 𝑈𝐺 𝜌𝐺 1/2 [𝑔𝜎(𝜌𝐿 − 𝜌𝐺 )]−4 (1)

The Kutateladze number represents the balancing of forces between inertial forces of the vapour,
surface tension and buoyancy and is particularly relevant when no characteristic dimension of the
testing apparatus can be determined. Pushkina and Sorokin suggested that flooding occurs when the
Kutateladze number attains a value of 3.2 numerically.

One of the first researchers to study the flooding phenomenon in terms of the Kutateladze number was
Richter (1981). Dukler and Maron’s (1984) research suggests three candidate mechanisms to
13

satisfactorily explain flooding - switching behaviour between multiple system steady-states that
induce upward flow of liquid, system velocities exceeding a kinematic wave's velocity of
propagation and entrainment phenomena. Switching behaviour between these states is
hypothesized to explain upward flowing film behaviour at the flooding onset.

McQuillan and Whalley (1985) compiled a database of flooding data, testing 17 empirical and 5
theoretical correlation equations to a total of 2762 data points amassed over 24 references,
demonstrating conclusively that non-dimensional correlations based on superficial velocities
deviate at high liquid flow rates. Critical flow regimes have been identified by Bezrodnyj and
Antoshko (1992) by measuring the film thickness acoustically. Comparison with past data helped
them identify at least some transitional boundaries satisfactorily.

3.2 Suggested Flooding Mechanisms


Prior research on flooding has substantial discrepancies, as researchers widely disagree on the
causal mechanisms leading to this behaviour. At low gas flowrates, the liquid flows preferentially
down the column walls, essentially creating a falling film. Flooding onset is characterized by the
appearance of wave-like patterns on the falling film. Progressive increase in the gas velocity leads
to film disintegration. In general, there exists no consensus among researchers regarding the
definitive causative factors for flooding.

Ring-like Wave Formation: Hewitt (1961) presents the first comprehensive analysis of flooding,
considering the onset of flooding to be accompanied by the appearance of large waves at the film
interface, below the liquid feed. Numerical instabilities are exhibited between the liquid film
velocity field and the observed gas flowrate at flooding, suggesting unstable wave formations as
a causative mechanism. Vijayan et al. (2001) however, investigates the effect of tube diameter on
flooding and puts forth that the wave formation mechanism is prevalent at smaller tube diameters
of 25 mm while droplet carryover or churn-motion in the liquid film prevails at higher diameter
of up to 99mm. Such effects have been observed previously by Suzuki (1977), McQuillan (1985)
and Govan et al. (1991).

Droplet Carryover: Vijayan et al. (2001) investigates flooding in tubes of 67-99 mm diameters and
proposes droplet carryover as the causative factor for large aperture tubes, subject to low liquid
flow rates. Under such circumstances, with increase in the vapour flow rate, the droplet flux is
augmented and gets eventually carried beyond the liquid entry point and is expelled out of the
system. However, the entrainment model predicts unreasonably large bubbles, hence, Maron
14

(1984) has suggested that this model does not provide a complete explanation to flooding, but may
only be a partial explanation.

Wave carryover via Churn-like Motion: Churn-like motion is the reason for flooding, at higher
diameter tubes (67mm), at sufficiently high liquid flow rates, as per Vijayan et al. (2001). Repeated
upward propagation and collapse of the large waves, aided with more droplet formation, gets
manifested macroscopically as a churning motion. Indirect observations of this effect have been
reported by Zabaras (1988) and Biage (1989). Research conducted by Mao (1993) presents
evidences that churn flow is but a manifestation of slug or plug flow. In a brief communication,
Hewitt (1993) responds to the above claim by explicitly distinguishing the two flows, yet admitting
that in some domains, these two flows essentially overlap. As a classical counter-argument, the
phenomenological models that are used to describe churn flow of the third kind differ distinctly
from the ones used to model annular flow, hence the distinction of these two flows is justifiable.
Hewitt and Jayanti go on to suggest that the two other categories of churn-flow may be better
described as churn-annular or churn-slug flow to eliminate any discrepancy and confusion. In a
brief communication, Jayanti et al. (1993) and researchers have conclusively observed Taylor
bubbles as definite proof for the occurrence of slug-flow which leads to the churn flow of the third
type, prior to flooding. Churn-flow has also been visually observed by Govan et al. (1990, 1991)
and modelled by Barbosa Jr. et al. (2001), employing a test section with a specially-constructed
transparent liquid inlet.

Fig 1: Schematic representation of the formation of a Taylor bubble prior to flooding (reproduced from Jayanti et
al., 1993)
15

Film Thickness Reduction: Dukler et al. (1984), suggests concentric waves to be the consequence,
and not the cause of flooding. Their research investigates the mean film behaviour of the interface,
which essentially begins to diminish, close to the flooding onset, for medium to high liquid
flowrates. By tracking the RMS properties as time series models, their research demonstrates that
a reduction of the mean film thickness is manifested as an increase in the fluctuations of the
pressure gradient and the interfacial shear stress.

Till date, researchers do not possess a commonly agreed upon criterion for flooding. Vast
disagreements exist among researchers when it comes to modelling such behaviour, and although
semi-empirical researches have suggested that to each of the above cases there is a possible cause,
no research exists till date which conclusively provides an explanation of the process, matching
experimental data conclusively. Not only are the causative mechanisms inherently complex, but
they are also interrelated. Chemical literature shows that the phenomena depends on several
mutually interacting parameters. Hence, it is the need of the hour to attempt to investigate the
behavioural dynamics of the process, if one is to come up with a suitable and sufficiently holistic
mathematical model that describes it satisfactorily.

3.3 Parameters Affecting the Flooding Onset

Extensive literature reviews clearly show that there are some very unique, discernible parameters
that clearly affect the onset of flooding. We attempt to briefly mention each such parameter:

Length Effects: The dependence of flooding with change in the pipe length has been explicitly
observed and recorded. Suzuki (1977), Hewitt et al. (1965), McQuillan (1985) and Govan et al.
(1991)'s experiments prove that not much can be conclusively said today about how the channel
length influences flooding. These researchers have employed a porous wall to introduce the
flowing liquid to the pipe. In fact, other than Suzuki (1977), all others have employed porous walls
to remove the liquid. Some other researchers have discharged the liquid film to the lower plenum
and have reported diametrically opposite results. Some reports suggest the flooding gas velocity
to decrease with increasing pipe length while others report that the pipe length is not a
determinant for the flooding phenomena.

Various results demonstrate what is known as the ‘length effect’. At the highest water rates, the
gas flooding diminishes by a factor of about 2 as the column length is increased. Wallis (1965)
confirms the same results as column lengths increase from 9 in. to 12 ft. Bankoff (1986) observes
that the effect of pipe length is more prominent at elevated liquid flow rates. Jeong (1996)’s results
16

conclude that the pipe length does not affect the flooding onset, provided a rough or a sharp end
geometry exists.

Entrance, Exit Effects and End Geometry: Entrance and exit effects play a significant role in
enhancing or diminishing the flooding behaviour. The best attempt to understand this is done by
Jeong et al. (1996). Their work suggests that provided there is a rough or a sharp-end geometry at
the edge, the flooding onset is not affected substantially. The research also shows that the end
geometry has significant contribution to aiding or hindering the flooding onset. Their results
indicate that the flooding onset is unchanged as long as the end has a sharp-edged geometry.
Contrastingly, the onset varies when tubes having a smooth exit and entrance is used. This
suggests that while sharp edges break down the propagating wave disturbances, smooth pipes
actually aid their growth. The end-geometry is an important parameter, but not much significant
research have been performed by past researchers before. For sharp exit geometries and low liquid
flow rates, the flooding onset and the instant where the flow disruption occurs, followed by liquid
expulsion, correspond to different temporal states of the multiphase system. Also, their work
proves that the effect of length is only visible when the end geometries are smooth.

Pipe Diameter: Vijayan et al. (2001) and co-workers have constructed an elaborate flow circuit,
attempting to understand this variation. Their results show that while wave formation occurs at
smaller aperture tubes, churn-flow like motion and droplet entrainment are more significant
causes that lead to flooding onset at higher tube diameters. They go on to explain how prior to
their work, researchers have only been able to partly explain their data, because they took into
consideration only one cause, rather than a combined action of these three causes, which switch
with increment in the tube diameter. The direct effect of pipe diameter is seen on the wall pressure
measurements via Kulite pressure transducers as studied by Choutapalli and Vierow (2010) for
the air-water system. They conclude that flooding may be caused by a constructive superposition
of waves rather than just one wave, as evidenced by the lack of a dominant frequency in the axial
location of the Fourier spectra.

Liquid Viscosity: The most comprehensive review on this parameter is described by Clift et al.
(1964). From intuition, a more viscous fluid has more resistance to reverse its flow condition.
Similarly, it is harder to reach the flooding onset owing to increased viscous resistance to wave
formation, droplet entrainments and churn-flow like motion. For less viscous fluids like water, the
Wallis-type correlations are followed satisfactorily. However, there is a marked deviation from
these empirical equations for more viscous fluids, pertaining to Reynolds number ranging from
100 to 1100. When a wetted wall column floods, liquid flow down the column immediately
17

decreases. This effect is seen to progressively lessen with increase in the absolute viscosity of the
fluid chosen.

In more recent years, Hlaing et al. (2007) and co-workers have investigated the viscosity effect on
such flows, with measurements primarily in the bubble-to-slug transition domain. Viscosity was
seen to lower the pressure gradients with increasing air Reynolds number for bubble, slug and
slug-churn flow regimes. However, pressure gradients are seen to diminish in the annular and
mist flow regimes. A comparison between theoretical and experimental pressures is made by the
researchers by using the equations predicted by Nicklin et al. (1962) for the bubble and slug flow
regimes and the Lockhart-Martinelli model equations in the annular and mist flow regimes.

Surface Tension: Deendarlianto et al. (2010) presents a fairly comprehensive treatment on this
aspect, focussing on inclined tubes, choosing diameters of 16mm and 1.1m. Surface tension seems
to significantly affect flooding at higher liquid flow velocities. With increase in the surface tension,
two-phase mixtures and even water is seen to exhibit tube blockage. Droplet entrainment
competes as a causal mechanism for flooding, at lower surface tension values and lower superficial
gas flooding velocities. However, it is clear that with more inclination in the tube, the Wallis
correlations progressively deviate from experiments.

Inclination angle of tube: Zapke's (2000) research is primarily focused on understanding the effect
that inclinations have in altering the various transition regimes associated with flooding. The
research highlights the limitations of some previous methods. Of particular focus is the fact that
the liquid Froude number and the Kutateladze number fail to account for inclinations and length
effects. A better correlation is obtained from the Ohnesorge number.

𝜇𝐿 2
𝑂ℎ𝐿 = √𝜌 (2)
𝐿 𝜎𝐷𝑒

In the second part of this revolutionary paper, the authors provide a much better reasoning of their
choice of the Ohnesorge number as a better correlation parameter. Their introduction of the
Froude-Ohnesorge correlation is consistent with the reported data of Clift et al. (1966) and Chung
et al. (1980). According to Zapke and Kroger (2000), the following is a better flooding correlation:

√𝐾𝑢𝐺 √𝐾𝑢𝐿
+ 𝑚 tanh[𝐶 = 𝐶1 (3)
tanh[𝐶2 (𝐵𝑜 1/8 )] 2 (𝐵𝑜 1/8 )]

The gas phase superficial velocity will change, if the tube is inclined, as per the following equation,

𝜌
𝑈𝐺∗ = 𝑈𝐺 √2𝑔𝐻𝑠𝑖𝑛𝜃(𝜌
𝐺
(4)
𝐿 −𝜌𝐺 )
18

Ousaka et al. (2006) has also studied the same phenomenon, and suggests an empirical correlation
to predict the gas flooding velocity, in terms of dimensionless numbers as,

𝐿 0.01
𝑅𝑒𝐺 = 1.35 ∗ 10−4 𝑅𝑒𝛤 −0.12 (𝐷) 𝑊𝑒 0.55 𝐹𝑟𝐿 −0.01 (5)

Liquid Film Thickness: A film model, put forth by Taitel and Barnea (1982) characterises the
flooding onset and flow reversal in terms of the liquid film thickness as the primary controlling
parameter. The film model requires an expression for the interfacial shear as input and predicts a
hysteresis between flooding and flow reversal. Results seem to be well in agreement with
experimental data. Optimal, physically realistic solutions are attained using linear stability
analysis and limiting flow as dictated by the kinematic wave theory of Lighthill and Whitham
(1955). Qualitatively, the film-theory follows experimental patterns, exhibiting a hysteresis loop at
higher liquid flow rates.

Karima and Kawaji (2000) have employed high speed video imaging and photo-chromatic tracer
dye techniques to understand the effect of liquid film on flooding. They propose that the flooding
onset is identified by interfacial waves not blocking the tube, and not propagating below with
appreciable velocities. Deendarlianto et al. (2005) has investigated liquid film behaviour as a
potential parameter that characterizes the flooding behaviour in multiphase systems via visual
observations, employing a constant electric current method. Their research sheds light on a new
aspect- that there are possibly two mechanisms of flooding. Lower flooding occurs at lower water
flow rates and higher tube inclination. Upper flooding occurs at higher water flow rates and lower
inclination angles. Lower flooding is seen at higher tube inclinations and lower water flowrates,
whereas upper flooding is seen to occur at lower inclinations of the test sections, and at higher
water flowrates. Liquid film characteristics in vertical narrow channels have also been studied and
correlated experimentally by Drosos et al. (2006).

Condensation Effect: Lee and Bankoff (1984) presented their findings of the parametric effects on
the flooding onset. Condensation effects in this paper are based on exhaustive top and bottom end
flooding experiments. Condensation effects have local as well as global implications. The local
properties of the formed interfacial waves vary for condensing flow, than say, compared to
adiabatic flows. The effects are more prominent in steam-water multiphase systems, typically
encountered in nuclear reactors. Bottom flooding is typically, less heat-transfer limited and is more
limited to the hydrodynamics of the encountered flow.

Interfacial shear stress: Preliminary investigations by Johnston (1984) led him to believe that the
interfacial shear is an important causal mechanism for flooding. Such semi-stratified or stratified
flows may be satisfactorily described in terms of semi-empirical equations, coupled with the
19

equations for momentum balance. Interfacial stresses are strongly influenced by surface tension
gradients, and viscous effects have significant influence on this. Guedes de Carvalho and Talaia
(1998) seem to agree. They extend their treatment to investigate slug-flow stabilities, and propose
𝜏𝑖𝐹
that the term indicates satisfactorily, the flooding onset, with moderate temporal variations.
√𝜌𝜎𝑔

3.4 Mathematical modelling of the Flooding and Post-Flooding Onset

Past researchers have attempted to characterize the onset of flooding by some mathematical, semi-
empirical correlations that describes this flow reversal condition. What is more, researchers even
widely disagree as to how they define the "onset of flooding". Sherwood et al. (1938) looks at it as
the point where there is violent bubbling of the gas through the liquid, accompanied by drastic
augmentation of the pressure drop, when there is an infinitesimally small increase in the gas
velocity. Wallis et al. in 1961, considered the flooding point as the point where the liquid begins to
exhibit chaotic dynamics. Working along with Hewitt et al. (1963), they characterized the flooding
state as a state when the liquid is chaotic, resulting in expulsion out of the tube.

Time-series models to understand the film behaviour: Salazar (1976) attempted to measure the
time-averaged film thickness by laser-scattering. While some sort of functional variation is visible,
it does not clearly explain the temporal variations and its relation to the film behaviour.

Dukler et al. (1984)'s research demonstrates that it is in fact, changes in the structural integrity of
the wall film is the precursor to the flood point, typically occurring at gas rates below it. What
adds credibility to their research is that the treatment is extensively mathematical, employing time
series and probability distribution functions as relevant statistical aids. Their interpretation of the
RMS film thickness is:

δRMS = [ { δ(t) - <δ> }2 ]1/2 (6)

<δ> represents the time-average film thickness for the investigated system. Prior to the approach
of the flooding point, there is another point called the loading point, which is characterised by
sudden increase in the δRMS value. A much drastic increase in the δRMS values shortly after indicates
the onset of flooding.
20

Lacy and Dukler (1994) put forward a new film model to predict the entry region flooding,
following exhaustive experiments. They propose that the wave mechanism is in fact, invalid. Their
research demonstrates that flooding occurs at the feed entry point, as validated satisfactorily by
mean film thickness measurements in the feed region.

Paglianti et al. (2000)’s work attempts to identify the onset of loading and its transition to flooding
employing a time-series analysis via a method involving non-intrusive conductance probes.
Quantitative statistical analysis highlights sharp differences between the two flow states, majorly
through the analysis of zero-level crossings employing conductance signals. Though the treatment
is done primarily for reactor tanks, the underlying mechanism may be clearly employed for
predicting flooding in continuous-contact multiphase equipments like tubes, absorption columns
and distillation columns.

Vijayan et al. (2001) and co-workers have seen flooding as a result of three inter-related forces in
action, namely:-

<i> Formation of ring-type waves that propagate upwards along the column length, causing
flow instabilities.
<ii> Entrained liquid droplets being carried over by the vapour flow rate and
<iii> Enhanced churning-like motion which carries over liquid via large waves.

While the first type is rather common to observe at low vapour flow rates, the latter mechanisms
only become prevalent at higher flow rates, typically for higher diameter channels. Each of these
three proposed mechanisms are in partial agreement with previously recorded results.

In all the past literature cited, researchers have used semi-empirical correlations, with attempts to
correlate the flow behaviour in such systems. The majority of these correlations aim to interrelate
either the non-dimensional superficial velocities or the non-dimensional Kutateladze numbers for
the liquid and gas phases. The most frequently used of these correlations now follow.

Wallis-type correlations: Correlations have been recorded by Wallis (1961) that correlate the
flooding gas and superficial liquid velocities. These velocities are interrelated as:

√𝑈𝐺∗ + 𝐶1 √𝑈𝐿∗ = 𝐶2 (7)


21

where 𝑈𝐺∗ and 𝑈𝐿∗designate the dimensionless, superficial vapour and liquid velocities, given as:
𝜌 𝜌
𝑈𝐺∗ = 𝑈𝐺 √𝑔𝐷∆𝜌
𝐺
and 𝑈𝐿∗ = 𝑈𝐿 √𝑔𝐷∆𝜌
𝐿
(8)

Experimental runs approximate the values of the constants as 𝐶1=1.0 and 𝐶2=0.88 for the air-water
system. These constants are dependent on the type of liquid being used, and therefore, will change
for the air-oil system, the focus of this present work. This is one of the earliest works in chemical
literature, interrelating the flow velocities of flooding, non-dimensionally. Wallis-type correlations
have the tube diameter as a flow determining parameter.

Alekseev-type Kutateladze correlations: The gas-phase Kutateladze number may be obtained via
the following equation, as used by Alekseev et al. (1972):

𝐾𝑢𝐺 = 0.2576𝐹𝑟 −0.22 𝐵𝑜−0.26 (9)

The Froude number (Fr) and the Bond number (Bo) are defined as follows.

𝐿𝑔0.25 (𝜌𝐿 −𝜌𝐺 )0.75 𝐷 2 (𝜌𝐿 −𝜌𝐺 )𝑔


𝐹𝑟 = 𝜎 0.75
and 𝐵𝑜 = 𝜎
(10)

Sun-type Kutateladze correlations: Building on the work of Wallis, Sun (1979), attempts to
correlate the flow dynamics in terms of the non-dimensional Kutateladze numbers of the two flow
phases. Their correlations are as follows:

∗1/4 ∗1/4
𝐾𝑢𝐺 + 𝐶3 𝐾𝑢𝐿 = 𝐶4 (11)

where, 𝐾𝑢𝐺∗ and 𝐾𝑢𝐿∗ are the dimensionless gas and liquid superficial velocities defined as:

1/4 1/4
𝜌2 𝜌2
𝐾𝑢𝐺∗ = 𝑈𝐺 (𝑔𝜎∆𝜌
𝐺
) and 𝐾𝑢𝐿∗ = 𝑈𝐿 (𝑔𝜎∆𝜌
𝐿
) (12)

For these set of correlations, 𝐶3 = 1.0 and 𝐶4 = 1.79, when tested for the air-water system.
This is a better correlation take into account not only the tube diameter, but also the surface tension.

Jayanti-Hewitt's correction to Wallis-type correlations: The Wallis-type correlations were better


fitted by Jayanti et al. (1996) by a slight modification of having an expression for C1 which is
dependent on the L/D ratio. Their research shows that for very sharp entry, the value of the two
constants change to 𝐶1=1.0 and 𝐶2=0.75. Hence, the choice of these constants is purely a matter of
choosing the appropriate flow regime, which is influenced by the L/D ratio for the system.

C1 = 0.1928 + 0.01089(L/D) - 3.754*10-5 (L/D)2 for L/D <= 120 (13)

C1 = 0.96 for L/D>120 (14)


22

where, L is the distance between the liquid inlet and the liquid outlet.

Post-flooding correlations: Vijayan et al.'s (2002) analysis of the post-flooding conditions for the air-
water system go on to predict fairly accurately, the various flow parameters (pressure gradient,
friction factor etc.) just prior to flooding onset. They conclude that while flooding onset are defined
fairly well from the correlations mentioned before, discrepancies exist for the pressure gradients.
The present work relates the friction factor and the pressure gradient to interfacial shear stresses.

𝐷 𝑑𝑃
(− )𝛼2.5
4 𝑑𝑧
𝑓𝑖 = 1 (15)
𝑈𝐺 2 𝜌𝐺
2

where, α is the post-flooding void fraction, and may be related to the film thickness as,

2𝛿 2
𝛼 = (1 − 𝐷
) (16)

The post-flooding film thickness depends on certain non-dimensional groups. Non-linear regression provides

an approximately valid correlation for this flow regime. Here 𝛿𝑡ℎ represents the dimensionless theoretical
film thickness and is obtained by Nusselt’s theory or from Fiend's (1960) correlation, which is
dependent on the film Reynolds number.

∗ )
𝛿𝑝𝑓 = 𝑓(𝑅𝑒𝛤 , 𝑅𝑒𝐺 , 𝐷 ∗ , 𝛿𝑡ℎ = 2.5 𝑅𝑒𝛤 −0.15 𝑅𝑒𝐺 0.1 𝐷 ∗ −0.3 𝛿𝑡ℎ

(17)

Fig 2: Qualitative profiles depicting the flooding behaviour with change in the vapour and liquid flow rates.
[Reproduced from http://thermopedia.com based on Mc.Quillan and Whalley (1984)]
23

Researchers however, agree that wave formations are one of the effects that is indicative of flooding.
However, the precise nature of the cause of flooding is still uncertain, primarily due to entrance effects.

The creation of a vena contracta causes the generation of sharp, unstable waves. Experiments demonstrate
that these waves are partly coherent at larger tube diameters, and entrainment at the wave tips occur
prior to the waves. These are transported upwards, and in turn, cause upward transport of liquids above
the tube. For larger diameter tubes, an alternative flooding criteria is proposed by Pushkina and Sorokin
(1969) in terms of the Kutateladze number.

1
1/2
𝐾𝑢𝐺 = 𝑈𝐺 𝜌𝐺 [𝑔𝜎(𝜌𝐿 − 𝜌𝐺 )]−4 =3.2 (18)

Hence, what is seen is that there are vast discrepancies in the criteria that researchers have set to define
the onset of flooding. There is hence, a need to establish a system that would sufficiently describe the
phenomena. Such a model would be consistent also with previously recorded experimental data. It
would be able to explicitly describe the phenomena satisfactorily yet sufficiently, consistent with the past
literature review of Chakraborty S (2016).

Flow reversal occurs for smaller diameter tubes,


and may be approximated by:

1
1/2
𝑈𝐺∗ = 𝑈𝐺 𝜌𝐺 [𝑔𝜎(𝜌𝐿 − 𝜌𝐺 )]−4 =1 (19)

For tube diameters exceeding 3 cm, Eqn. (18)


becomes a better mathematical model.

Fig 3: Comparison of flooding data for sharp-edged, bell-


shaped (taped) and porous wall liquid exit [reproduced
from Govan et al. (1991)].

3.5 Mathematical Characterisation of Flow Regimes in Flooding

Anderson and Mantzouranis (1960) investigated pressure drop and holdup effects for two-phase
flow systems, in the annular regime. Their treatment is more classical, in the context that they
employ the universal velocity distribution of turbulent flow to understand the flow behaviour.
Hence the method relates the wall shear stress to the flow velocity. We define the flow velocities
in terms of the friction velocity 𝑢̈ ∗ :
24

𝜏𝑤
𝑢̈ ∗ = √ (20)
𝜌

whence, we get 𝑢+ and 𝑦 + as follows.

𝑢 𝑦𝑢̈ ∗ 𝜌
𝑢+ = 𝑎𝑛𝑑 𝑦 + = (21)
𝑢̈ ∗ 𝜇

The treatment segments the flow domain into three sub domains- the laminar sub layer, the
intermediate buffer layer and the turbulent core. Hence, we have:

𝐿𝑎𝑚𝑖𝑛𝑎𝑟 𝑆𝑢𝑏𝑙𝑎𝑦𝑒𝑟: 𝐹𝑜𝑟 𝑦 + < 5, 𝑢+ = 𝑦 +

𝐵𝑢𝑓𝑓𝑒𝑟 𝐿𝑎𝑦𝑒𝑟: 𝐹𝑜𝑟 5 < 𝑦 + < 30, 𝑢+ = −3.05 + 5 𝑙𝑛𝑦 +

𝑇𝑢𝑟𝑏𝑢𝑙𝑒𝑛𝑡 𝐶𝑜𝑟𝑒: 𝐹𝑜𝑟 𝑦 + > 30, 𝑢+ = 5.50 + 2.5 𝑙𝑛𝑦 + (22)

The turbulence model act as a cause for flooding, particularly at low values of the gas phase
Reynolds number. However, this is not valid once the flow becomes annular.

A unique treatment was made by Bharathan and Wallis (1983) to understand the flow regimes in
air-water counter current flows. Their work identifies the flooding phenomena to occur in four
stages. The first stage is a smooth film, where the wall shear stress is much larger than the
interfacial shear. (τw>>τi) The liquid flows undisturbed through the pipe length as a continuous
film. Nevertheless, the gas flux, at any instant, limits the downward flow of liquid. This is a regime
of smooth film. Transition occurs to a rough film upon interfacial agitation. More liquid enters the
tube than can exit and a thicker film is generated. The authors propose that in order to maintain
an equilibrium film thickness, liquid is carried up by the gas. Eventually, the film gives way to the
dry tube, just prior to the flooding onset. Each of these regimes may be analysed now.

Smooth Film (Regime A): This regime is manifested as a smooth interface, and the flow limiting
mechanism for the liquid counter-flux occurs at the top end geometry. A nearly free-falling film is
seen and the liquid flow rate is majorly governed by the wall shear stress. Viscous effects in the
film are in terms of the liquid Reynolds number:

𝜌𝐿 𝐷
𝑅𝑒𝛤 = 𝑢𝐿 (
𝜇𝐿
) = 𝑢𝐿 ∗ 𝐺𝑟𝐿 (23)

where 𝐺𝑟𝐿 is typically, a form of the Grashof number and 𝜂𝐿 is the fluid's kinematic
viscosity.

√(𝜌𝐿 (𝜌𝐿 −𝜌𝐺 )𝑔𝐷)𝐷


𝐺𝑟𝐿 = 𝜂𝐿
(24)
25

Wallis (1969) has suggested that for 𝑅𝑒𝛤 <1000, the flow remains laminar. Under such conditions,
the film thickness is related to the pipe diameter as

𝛿 3𝑢 ∗ 1/3
= (4𝐺𝑟𝐿 ) (25)
𝐷 𝐿

For turbulent flow, based on the frictional wall effects, Wallis gives the empirical equation,

𝛿
𝐷
= 0.063𝑢𝐿 ∗ 2/3 (26)

Rough Film (Regime B): Large interfacial shear significantly agitates the liquid film over the gas
core. Since the flow in either phase is turbulent, characterized by wavy interfaces, the friction factor
may be best related to the film thickness empirically as a constitutive relation of the form,

𝑓𝑖 = 0.005 + 𝜋(𝛿, 𝐷, 𝜎) (27)

For large tubes, the Kutateladze number, as proposed by Pushkina and Sorokin (1969), governs
the onset of liquid film penetration during flooding.

𝐾𝑢𝐺 = 𝑢𝐿 ∗ √𝐷∗ = 3.2 (28)

The pressure gradient is seen to vary with the liquid fraction as per the following relationship.

𝑑𝑃 ∗
− ( 𝑑𝑧 ) ≈ (1 − 𝑧) (29)

Discontinuity (Regime C): When a film discontinuity appears, counter flow gets restricted at both
ends of the tube. The flow is typically unsteady, dynamic and oscillates spatially along the tube
length. The amplitude of oscillations naturally, are strong functions of the input line impedance
and flow pattern and characteristics.

Flooding envelope (Regime D): In this stage, both the interfacial wall stress and the wall shear
stress become dominant, to yield a flow devoid of any abrupt discontinuity in the film. Ideal
flooding conditions corresponding to a flooding envelope are hard to create due to design entry,
exit, tube geometries and the variety of flow conditions.

However, based on the work conducted by researchers all over the world, there are some distinct
flow regimes that are observed, prior to flooding. These are identified as follows:

Annular Flow Regime: Wolf et al. (2001) has investigated the behaviour in annular flows, leading
to flow creation in air-water systems. Their results prove that the first regime, viz. annular flow,
takes some time to develop, typically about 100 pipe diameters. Changes that follow are primarily
26

due to variations in the gas densities, due to pressure drop. Film flow rate and the film thickness
are the most insensitive variables for this state.

Wilkes et al. (1983) reports wave coalescence and entrainment as visually perceivable effects for
annular flow. Two types of waves, namely minor ripples and disturbance waves, cause the transfer
of mass and momentum in such systems. The Monte Carlo method is used; the resulting treatment
is probabilistic, the random numbers being generated from the Harwell Subroutine Library
program FA03A. The initial wave distribution, is assumed to be a cut-off normal distribution.

√2 −(𝑟−𝜆)2
𝑝(𝑣) = 𝜆
𝑒𝑥𝑝 (2𝑠𝑑 2
) (30)
𝑠𝑑√𝜋(1+𝑒𝑟𝑓{ })
𝑠𝑑√2

The earliest attempts by Kirillov et al. (1987), attempted to understand this flow regime
mathematically by presenting a two-dimensional mathematical model for annular and dispersed-
annular flows. Though the model is generic to two-phase flows and is not focused on flooding,
their approach and the underlying mathematical rigour in the presented model is certainly worthy
of mention. Though they consider a vertical tube, the difference lies in the fact that they incorporate
electrical heating to add heat-transfer effects, which leads to the creation of two-phase flow.

Govan et al. (1989) has successfully measured the shear stresses and its temporal variation for such
flows employing heat transfer methods. This was majorly because they find that most
measurements of shear stresses are indirect, based on analogies between momentum and heat
transfer (Reynolds analogy) or momentum and mass (Chilton-Coulbern's J-factor analogy).

Slug Flow Regime: In a brief communication, Taitel and Dukler (1976) attempt to present a new
theoretical approach to the correlations of Lockhart and Martinelli (1949) for stratified flows. The
solutions for the liquid holdup and pressure drop are convergent with the Lockhart-Martinelli
correlations. The treatment is generic to stratified flow and does not address any specific flow sub
domains. Other notable attempts have been done by Johannessen (1972) and Whalley (1987).

Taitel and Dukler (1977) characterize this regime as one associated with coherent plugs of aerated
liquid, separated by zones of gas riding on the thin liquid film. What is noteworthy is that in this
domain, Blasius type correlations give better initial estimates and better convergence for the
friction factors (which may be evaluated by equations like the Colebrook-White type). Momentum
and mass conservation equations are solved using the finite-difference scheme to predict profiles
in good agreement with data. A generalized non-dimensional fundamental model is presented
that predicts slug flow fairly accurately for systems subject to entry based natural slugging.
27

Churn Flow Regime: Govan et al. (1991) investigates churn flow behaviour in vertical pipes.
Dukler (1984) has investigated and proposed a physical model for prediction of the minimum slug
length 𝑙𝑠 . Using boundary layer assumptions, the proposed model attempts to understand
physical mechanisms pertaining to slug length stabilisation. In fact, the recurrent relaxation and
development of wall boundary layers are useful in detecting the minimum stable slug length.

3.6 Surfactants

Surfactants are organic amphiphilic compounds, and comprise of a hydrocarbon hydrophobic tail
and a polar group acting as the hydrophilic head [34]. They are commonly classified on the basis
of the ionic characteristics of the polar head group, into anionic, cationic, non-ionic or zwitter-ionic
(Ottewill, 1984). The hydrophile-lipophile balance (HLB) is commonly used internationally to
categorize surfactants. This number indicates the relative tendency of a surfactant to dissolve in
water and hence, form oil-in-water or water-in-oil emulsions. A surfactant that is more soluble in
oil and forms water-in-oil emulsions is assigned a low HLB number. A surfactant that has a low
HLB number should be selected when the formation salinity is low. Such surfactants exhibit micro-
emulsion at low salinity levels. Griffin suggests an equation for non-ionic surfactants to evaluate
the HLB:

HLB = 20MWh / MW (31)

where MWh is the molecular mass of the hydrophilic portion of the molecule, and MW is the
molecular mass of the whole molecule, giving a result on an arbitrary scale of 0 to 20.

An HLB value of 0 indicates a completely hydrophobic molecule while an HLB value of 20


corresponds to a molecule made up completely of hydrophilic constituents.

● HLB value between 0 to 3 indicates an antifoaming agent.

● HLB value between 4 to 6 indicates a water/oil emulsifier.

● HLB value between 7 to 9 indicates a wetting agent.

● HLB value between 8 to 18 indicates an oil/water emulsifier. .

● HLB value between 10 to 18 indicates a solubilizer or hydrotrope.

● HLB value between 13 to 15 is typical for common detergents

Davies suggested an alternative route to evaluate HLB, taking into account the effect of stronger
and weaker hydrophilic groups.
28

HLB = 7+ mHh - nHl (32)

In this equation, m is the number of hydrophilic groups in the molecule, H h is the value of the
hydrophilic groups, n is the number of lipophilic groups in the molecule and Hl is the value of the
lipophilic groups.

Sulphonated hydrocarbons are most commonly employed in surfactant flooding, leading to the
formation of petroleum sulphonates. This is because, above 200 0C, sulphonate groups can
maintain their thermal stability, while other surfactants rapidly decompose. The thermal stability
of sulphonates, as investigated (Zeigler, 1988), follows the trend below:

petroleum sulphonates < alpha olefin sulphonates < alkyl aryl sulphonates

Sulphate based surfactants, have in general, a greater tolerance to divalent ions, and
correspondingly, and are subject to hydrolysis at high temperatures and low pH, owing to their
ester linkages. Surfactants are hence, flow-enhancing additives which reduce the wall friction and
the interfacial surface tension. Investigations by Farn (2006) prove that surfactants can act as
foaming or wetting agents. For low pressure-drop systems, Liu et al. (2014) analyzes the liquid
holdup, pressure drop and drag force reduction and concludes that bubbly, churn and annular
mist-flow patterns are the most widely observed flow behaviour in these systems, post surfactant
addition. Surfactant addition induces a maximum holdup of 88.6%, in the churn-flow regime.
Additionally, the pressure drop reduction is up to 96.5% upon surfactant addition, which is
exhibited in the churn flow regime. While the two phase flow pattern is unaffected by the
incorporation of surfactants, the detailed flow configurations are significantly changed.

SDS or Sodium Dodecyl Sulphate is the most commonly used surfactant for investigating such flow
behaviour. With respect to air-water system, which is the present case, previous researchers have
attempted to characterise the surfactant addition in terms of whether the resulting flow behaviour
is foaming or non-foaming. Here, the CMC represents the Critical Micelle Concentration. From
Table 1, it is clear that most surfactants, upon addition, create foam in the multiphase system. We
next investigate how the addition of surfactants influence the flow mechanisms for multiphase
systems.
29

Article/ Journal Surfactant name CMC (ppm) Characteristics

(Sawai, Kaji, & Urago, 2004, pág. non Foaming/non-


CTCA / SO
Sawa) specified foaming

non Foaming
(Rozenblit , et al. 2006) DRA
specified

(Duangprasert, et al. 2007) SDS 2750 Foaming

non Non-Foaming
(Xia and Chai 2012) SDS
specified

(van Nimwegen, Portela and Henkes Trifoam 820 Block /


500 / 1700 Foaming
2013) SDS

(Liu, et al. 2014) HY-3 1000 Foaming

(van Nimwegen, Portela and Henkes Foaming


Trifoam 820 Block 500
2015)

Table 1: Characterization of Flow behaviour in air-water multiphase systems, subject to surfactant addition

3.7 Foaming in Surfactants

The phenomenon of foaming is a direct consequence of the use of surfactants in multiphase flows.
Foaming is the creation of trapped air bubbles, created by the complex hydrodynamic interaction of air
and water. Consequently, foaming significantly alters the flow dynamics and the rheological morphology
of the system. Foam results from the localised entrainment of air in the water phase and changes the
interfacial flow dynamics significantly. Foams are not stable in inherently pure liquids, they are stabilised
by the addition of surfactants, which is thought to alter the pressure drop, particularly at high liquid flow
rates. Guo-dong (2012) concludes that the addition of surfactant, is best manifested physically, as a
substantial decrease in the pressure drop, which is more prevalent in annular and stratified wavy flow
regimes. In fact, in the annular flow regime, the pressure gradient profile for the surfactant laden liquid
case, eventually converges to the pressure gradient profile for the pure-liquid case. For upward inclined
pipes, pressure gradients are seen to increase sharply, especially with increase in the superficial gas
velocity. Thus, adding surfactants would reduce the pressure gradient significantly, particularly in the
slug and annular flow regimes.

van Nimwegen et al. (2015) investigated the physical morphology of the air-water multiphase system,
upon the addition of surfactants. As mentioned by Khosla (2012), foam formation reduces the effective
density of the liquid, in turn postponing the transition from annular to churn flow at low values of the
superficial gas velocity. Additionally, foaming influences the gas-liquid interface morphology, which
30

influences the annular to churn flow transition. It has been suggested that surfactant addition changes
the static and dynamic surface tensions to change, generating a stable foam for such multiphase systems.

The effect of surfactant depletion on foaming has been investigated by Boos et al. (2012). Drainage, bubble
coarsening and bubble coalescence are the parameters influencing the foam-lifetimes. The numerical
quantification technique proposed for bubble characterisation is to employ the Sauter mean radius rather
than the arithmetic mean radius. Otherwise, even for moderate monodispersity indices of around 0.3,
significant discrepancies are exhibited between the calculated and measured depletion values if the
arithmetic mean radius is employed. The Sauter radius 𝑟32 and the polydispersity index PI are defined
as:

〈𝑟 3 〉
𝑟32 = 〈𝑟 2 〉 (33)

√〈𝑟 2 〉−〈𝑟〉2
𝑃𝐼 = 〈𝑟〉
(34)

Investigation into foaming phenomena indicates that surfactants laden water lessens the interfacial
tension and the wettability gradient, mostly due to the Gibbs-Marangoni effect. The addition of
surfactants lead to ‘stress-levelling’ which means that interfacial shear stresses are reduced, due to
increased elasticity of the interface. Under such circumstances, waves are less likely to attain large
amplitudes and propagate up the test section.

3.8 Effect of Surfactants on the Flow Morphology and Pressure Drop

Experiments have been carried out by researchers to understand and model this aspect of the
study. Their relative progression may be tracked in the following table. All researchers have
employed essentially a co-current system to model the air-water system, subject to surfactant
condition. The present research will however employ counter-current, surfactant loaded flows, as
would be described in detail later. The different experimental conditions that have been employed
by various researchers are detailed below:

Rozenblit et al. (2006) has investigated the effect of surfactants or DRA's (Drag Reducing Agents) on the
pressure drop reduction in tubes and channels. They employed a 52% surfactant, 48% water mix in
deionised water, subject to gentle stirring for a day. The addition of a surfactant is seen to significantly
reduce bubble coalescence. Also, the shapes of these bubbles are more spherical, than when pure water
is used. A certain oscillatory component of the pressure drop is augmented under bubble-flow
conditions; contrastingly for annular flows, this component decreases. The effect in the slug-flow
regime is marked. However, the research proves that surfactant addition lowers the net system pressure
31

drop, which gets more prominent in regimes of low superficial liquid velocities. Additionally, they find
the heat transfer coefficient is decreased in the presence of surfactants, more so at slug-churn transitions.

Inner pipe Liquid


Pipe length Gas velocity
Article diameter Inclination velocity
(m) range (m/s)
(mm) range (mm/s)
(Sawai, Kaji and
25.8 5.09 Vertical 0.05 to 15 2 to 800
Urago 2004)
(Rozenblit , et al.
24.2 20 Vertical 𝑢𝐺 /𝑢𝐿 is 64 Non-specified
2006)

(Duangprasert, et
19 3 Vertical 0.0026 to 58.81 0 to 123
al. 2007)

(Xia and Chai Horizontal,


59 11 0.07 to 50 20 to 1200
2012) 2.5°, 5°, 10°

(van Nimwegen,
Portela and 50 12 Vertical 21.5 to 43 2 to 40
Henkes 2013)

Non- 𝑢𝐺 /𝑢𝐿 is 100 to


(Liu, et al. 2014) Non-specified Vertical Non-specified
specified 10000

(van Nimwegen,
Portela and 50 12 Vertical 6 to 43 2 to 40
Henkes 2015)

(van Nimwegen,
Portela and 50 12 Vertical 21.5 to 43 2 to 40
Henkes 2015)

Table 2: Comparative study of the various set-ups established by various researchers to understand flow behaviour in the
presence of surfactants

Duangprasert et al. (2010) and co-workers have selected SDS as the surfactant and concluded via a series
of experiments, that the addition of surfactant lowers the critical gas phase Reynolds numbers for bubble-
slug flow and slug flow transitions. Bubble sizes increase and are more frequent at high Reynolds
number, in the presence of SDS than for the pure water case, which is attributed to lower surface tension.
Surfactant addition leads to the creation of a special class of Marangoni flows which leads to increased
surface elasticity. Interfacial rigidification effects are seen on surfactant addition which increases the
liquid viscosity, leading to reduced drop formation at low gas phase Reynold’s number. Confirming
Rozenblit's (2006) observations, pressure gradient diminishes in the presence of surfactants, which is
32

attributed to viscous interactions. Pressure gradient fluctuations are particularly noticeable in the slug
and churn flow regimes and is reflected in both the pure water and the SDS-laden water case. Also, their
research confirms the fact that the viscosity affects the flow pattern transitions.

Guo-dong et al. (2012) investigates the effect that surfactants have on the pressure drop for inclined tubes
and concludes that for all changes in the flow pattern, a corresponding change occurs in the pressure
gradient. Flow regimes (which were identified by electrical conductive tomographic technique) at four
inclination angles of 00, 2.50, 50 and 100 were analysed, upon which they concluded that the pressure
gradient decreases with increase in surfactants in the system, due to the Gibbs-Marangoni effect, leading
to surface tension reduction. The surface tension decreases with increase in surfactant loading in the
liquid phase. This reduction in the surface tension, is more pronounced for 95% purity grade SDS at low
concentration.

Significant improvements on this area has been done by van Nimwegen et al. (2013). Selecting two
surfactants (SDS and Trifoam 820 Block), both reduce the pressure gradient at low gas flow rates and
increase at high gas flow rates. Annular to churn flow occurs at the Turner criterion, which is commonly
used to predict the gas velocity, at the onset of liquid loading, when there occurs a reversal of the largest
droplets in the flow.

40𝜎𝑔(𝜌𝐿 −𝜌𝐺 ) 1/4


𝑢𝑐 = ( 𝜌𝐺 𝐶𝐷
) (35)

Here, 𝐶𝐷 denotes the drag coefficient, which is assumed to be 0.44 for very large droplets. They assume
the largest droplet size to correspond to a Weber number of 30. Thus,

𝑢𝐺 2 𝜌 𝐺 𝐷
𝑊𝑒 = ( 𝜎
) = 30 (36)

Different superficial gas velocities are used to check the validity and reproducibility of the results. The
researchers deduce a very interesting result. There seems to exist an optimal surfactant concentration at
every gas velocity, which minimises the pressure drop. However, at smaller concentrations, it is the
churn-flow that increases the pressure drop, while at higher concentrations, it is the intensified foaming
in the system that increases it. Also, the CMC is not a very good indicator of the surfactant dosage
required to suppress the pressure drop.

Further investigations are continued by van Nimwegen et al. (2015) on the air-water surface morphology
under surfactant laden conditions. They put forth the argument that for annular flow, foam generation
is suppressed by roll waves, creating a foamy crest on the ripple waves.
33

Fig 4: Flow regimes without and with Trifoam 800-Block surfactant, at 1000 wppm. [reproduced from van Nimwegen et
al. (2015)]

Fig 5: Influence of surfactants on the Air-Water System flow and Morphology [reproduced from van Nimwegen
et al., (2015)]

Van Nimwegen et al. (2013, 2015) and co-workers have identified distinct flow regimes on co-current,
surfactant laden air-water systems. Overall, the purpose of these works is to highlight the complex
morphology and the associated regime transitions in the air-water system, for co-current flows, with and
without surfactants. They find that foam being less dense than water, significantly alters gravitational
and inertial forces. In Fig 5, each flow conditions are contrasted with and without surfactant Trifoam 820
Block. The images depict the traversal of the liquid film (and droplets and ligaments close to it), with
upwards flowing gas. Foam forms through the entrainment of air in water, which is aided by the
34

formation of ripple and roll waves. At high gas and liquid flow rates, these ripple waves are markedly
diminished and roll waves are seen, which significantly makes the flow morphology even more complex,
while increasing the foam content. At low gas flowrates, the liquid film essentially exhibits churn-like
flow, which is almost independent of the liquid flowrate. This churn motion is characteristic of liquid
agitation, resulting in the formation of ligaments, droplets and bubbles at the interface. Foam formation
is aided additionally due to the flow being co-current; this foam results in the creation of foam waves
which transfers liquid upwards, the frequency of which is dependent on the liquid flowrates for the
system. For churn-like flow, the downward propagation of the liquid film and the consequent flooding
waves, are suppressed by the foam. Consequently, the foam tends to diminish the superficial gas velocity
at the transition of the flow morphology from annular to churn flow. A comparative analysis leads them
to conclude that foam formation plays the most crucial role in determining the multiphase flow dynamics
for air-water systems. The formation of foam leads to a change in the interfacial flow morphology,
suppressing the entrainment of large liquid droplets in the gas bulk. Referring to the previous
publications by the same team, it is evident that similar results are observed over a wide variety of water-
soluble surfactants. The researchers conclude that similar trends are exhibited and propose that this
qualitative and quantitative database be utilised in future to propose a mechanistic flow model that
explains the flow regimes and their associated regime transitions satisfactorily.

The combined standard uncertainty of the frictional pressure drop, as seen experimentally, is seen to fit
well with the following equation proposed by Liu et al. (2014).

2 2 1/2
∆ 𝑃𝑓 (𝑢𝐴 ∆𝑃)+(𝑢𝐵 ∆𝑃) 2
𝑢𝑐 (
𝐿
) = [( 𝐿2
) + (𝑔(𝜌𝐿 − 𝜌𝐺 )) + (𝑢𝐴2 (𝐻𝐿 ) + 𝑢𝐵2 (𝐻𝐿 ))] (37)

∆ 𝑃𝑓
where 𝑢𝑐 ( 𝐿
) represents the combined standard uncertainty of the frictional pressure drop per unit

length of the tube, 𝑢𝐴 ∆𝑃 and 𝑢𝐵 ∆𝑃 are terms representing the standard uncertainties obtained via Type
A and Type B kind of pressure drop evaluations and 𝑢𝐴 (𝐻𝐿 ) and 𝑢𝐵 (𝐻𝐿 ) represent the uncertainties
resulting from liquid holdup for Type A and B evaluations respectively.

The expanded uncertainty is obtained by multiplying Eqn. (19) with a coverage factor k, given by,

∆ 𝑃𝑓
U = 𝑘𝑢𝑐 ( 𝐿
) (38)

where U denotes the expanded uncertainty, and k, the coverage factor.

While there is a general consensus among researchers about the general observable effects of surfactant
addition, disagreements prevail when one takes a more Lagrangian approach to the system behaviour.
The exact effects of surfactant addition have not been satisfactorily described till date.
35

3.9 Future Research Scope: Possible Areas of Study

Based on this exhaustive review, one can clearly see the present limitations in the topic. No one has been,
till date been able to properly understand the length effect and the inclination effects on the flooding
onset. The mathematical approaches adopted by past researchers have been non-unifying and deal with
the problem in radically different routes. The multiphase dynamics of the air-water system have not been
completed understood. The parameters affecting the phenomena, and how the parametric variation(s)
influence flooding is not completely understood. Additionally, the incorporation of surfactants into the
system induce more complexity into the flow morphology and behavioural dynamics. Very little work
has been done for the air-water system with surfactants added to alter the flow characteristics. One
possible aspect that many researchers are attempting to investigate the effect of Taylor-Raleigh instability
on surfactant based multiphase system. A Taylor-Rayleigh instability is an instability that occurs at the
interface of two fluids of different densities when the lighter fluid exerts a push on the heavier one. This
effect is more prevalent for oil-water systems than for air-water systems.

Surfactants change the molecular interactions between the liquid and the gas phase. Consequently,
researchers have tried to come up with methods to augment surfactant performance in such multiphase
flow systems. Referring to the work of Zargertalebi et al. (2015), silica nanoparticles are employed to
increase the efficiency of anionic surfactants like SDS. It is concluded that such surfactants lead to better
enhanced oil-recovery. This is because, the surfactants are better able to capture and control the causative
mechanisms leading to flooding. AEROSIL 300 and AEROSIL R816 are the two nanoparticles used in this
study. Interfacial surface tension measurements (in accordance with the Young-Laplace's equation)
rapidly diminish at low surfactant loading, however for higher loading, the surface tension increases.

Yet another problem is the issue of scalability. While the air-water multiphase system is just a
representative, functional prototype, the eventual aim is to scale up these effects to commercial,
industrially utilized continuous-contact equipments like distillation columns. Scale-up, even today,
serves to be one of the most challenging domains in chemical engineering, because of the complex inter-
coupling mechanisms that may become even more dominant as one proceeds from a Lagrangian to an
Eulerian reference frame. While proportionality of the non-dimensional numbers is just a guideline
criteria, it is by no means an exhaustive route for scaling up a chemical process. It is only by conducting
more rigorous and target-oriented research that this problem may be solved. Only when we have
completely understood the air-water system (with and without surfactant addition) will we be able to
successfully map the results for the oil-water system. However, regardless of the system, flooding
remains a unique, mathematically-rigorous and physically relevant sub-domain of research in fluid
mechanics.
36

4 Project Objectives

In this SHELL funded project, the focus will be to investigate the dynamics of counter-current liquid and
gas phase flows. The project, under the Transient Multiphase Flow (TMF) Consortium, will attempt to
carry out experimental studies via axial photographic imaging of the flow, investigating the flow
behaviour with and without surfactants in the air-oil multiphase system. In a first-of-its-kind method,
analysis will be done to investigate how the incorporation of surfactant affects the interfacial dynamics,
wave characteristic and patterns, bubble growth, droplet entrainment, foaming, and the direct
implication of these effects on the flow regimes.

By a series of design experiments, the project will aim to understand the underlying morphologies, flow
dynamics and morphological behaviour of the oil-air multiphase system, with and without addition of
surfactants. The direct result of this project would lead to clear suggestions as to the types of surfactant
to be used, and in what concentration, in industrial oil and gas applications. The study will also be
beneficial in identifying the possible cause and effects of surfactant addition, and its advantages and
limitations, so as to assure superior flow quality assurance and to predict flow behaviours consistent with
experimental findings.

Preliminary documents
Preliminary lab work
Lab data recording
Data processing/analysis
Write-up
Milestone

Week 10

Week 11
Week 1

Week 2

Week 3

Week 4

Week 5

Week 6

Week 7

Week 8

Week 9

Safety tests
Set preliminary research objectives
Lab training and Rig deconstruction
Rig re-construction
Rig start-up: Flow lines tested and Leaks identified
Updating equipment: camera stand
Equipment testing: ready to record
Lower test section: oil-air measurements
Data analysis: Flooding onset
Data analysis: flow regimes and videos processing
Data analysis: videos (phenomenology)
Report write up
Pster Presentation
Thesis submission

Fig 6: Gantt chart showing the expected work plan leading to completion of the MSc. thesis
37

5 Experimental Methodology

5.1 Rig Revamping

The last runs on the Shell upflow rig were performed using Triton X-100, which is corrosive on the test
section material (Perspex). The surfactant creates fine, microstructural cracks and fissures in the entire
experimental setup including the upper and lower test sections, the injector and the extractor.
Consequently, none of the above mentioned parts were in operation when the research was to
commence.

As a first step, the entire rig was taken down and broken down into member components. All bits of
the test section were discarded as they had developed cracks. The injector and extractor were also found
to be corroded and were discarded. Some hoses were leaking in parts, while some hoses had surfactant
deposition on them. Flanges, which had been initially glued to the test sections, injector and extractor
were also cracked and corroded and had to be discarded. To surmise, no components of the rig were in
a condition good enough to be used. The entire rig was built from scratch. As a first step, the dimensions
of the flanges and test sections were determined and sent to the Chemical Engineering Workshop at
Imperial College London for manufacture. Following this, a new extractor and injector were fabricated
and attached to the rig. The manufacturing process taking about 2 weeks, support profiles were cut
from aluminium bars during that time, and attached to create a strong, vertical framework to
mechanically support the rig. Prior to installation in the rig, each of these profiles were machined to the
desired standards, filed and fines removed by using a pressurised air-gun. Hoses pertaining to the
appropriate dimensions were purchased from RS Components UK. Rotameters pertaining to different
flowrates, ranging from very low (0.3 LPM) to very high (40 LPM) were sourced to ensure widespread
variation in the liquid flow rates.

Once all the building components were available, the rig was rebuilt sequentially. As a new surfactant
was to be tested, the flanges were not glued to the test sections, rather they were optimally designed to
fit in an essentially leak-proof fashion. O-rings were added between each of the connections of the test
section, injector and extractor to prevent any residual leakages which might eventually be exacerbated.
Following the successful installation of the experimental apparatus, the liquid and gas lines were
individually leak-tested successfully. As creating an exhaustive flooding database was a target,
including (but not restricted to) very high liquid and gas flow rates, ring supports were specially
designed for the extractor and the injector, which were then successfully incorporated in the
experimental setup. At each stage of the process, the rig was regularly checked to ensure that it was
essentially vertical and that there were no inclinations to the column. A special, 10 m long, 180mm bend
radius reinforced hosing was ordered from RS Components UK to attach at the top of the rig, which
would essentially carry the oil-gas mixture at very high velocities, and therefore had to withstand the
38

appropriate flow duties. Once these components were successfully installed, the equipment was
successfully revamped and was now ready to be tested for the oil-air multiphase system, first without
surfactant and then, at different surfactant loadings.

It must be mentioned here that a digitally controlled, electronic mass flowmeter was employed in the
experimental setup to have maximum control on the gas velocity. Three specific softwares namely
FlowDDE, FlowView and FlowPlot, all licensed by Bronkhorst UK Ltd. were employed to control the
flow valve on the mass flowmeter electronically, which in turn controls the gas flowrate. Post the rig
setup, a high-speed camera and a halogen lamp were also installed on the rig to perform axial imaging
on the test section. The Imperial “i-speed 3” software suite is used to trigger the camera electronically,
which then captures images. The camera being used simultaneously for two other experimental setups,
a more focused scope of research had to be identified, one that would be possible to achieve the desired
results despite limited time constraints. Another problem that presented itself during the latter part of
the project was the fact that the surfactant Baker Petrolite tends to attack the material of the O-ring:
over time the injector started leaking. This problem was immediately tackled with by detaching the
flange between the injector and the upper test section and replacing the O-ring with a fresh one. In
retrospect, the injector would have been perhaps deemed useless, had the flanges been glued together
beforehand, as the nature of the leak is a chemical interaction between the surfactant and the O-ring
material, information that was unavailable a priory.

5.2 Materials

Exxsol-D80 is the light oil which was used in the present setup. This oil was sourced from ExxonMobil
Ltd. Flow parameters for this oil may be easily sourced online. Baker Petrolite is used as the surfactant
and is one of the Petrolite copolymers widely manufactured by Baker Hughes. This is an industrial
defoamer (and hence, foaming is not seen in the interfacial topology), that is stable under ordinary flow
conditions. It has a CMC value of 120 wppm. The flow characteristics are investigated at loadings
higher than its CMC (viz. 500 wppm and 1000 wppm. The surfactant has the following characteristics.

35
Defoamer concentration and
Surface tension ɣ [mN/m]

30
25 surface tension seem to have an
20 inverse relationship to each
15
other.
10
5 Fig 7: Reduction in the surface
0 tension with increase in the defoamer
0 1000 2000 3000 4000 5000
concentration of Baker Petrolite
Defoamer concentration c [wppm]
39

35
30
Surface tension, ɣ [mN/m]

25
0 wppm
20 25 wppm
15 100 wppm

10 250 wppm
1000 wppm
5
5000 wppm
0
0 500 1000 1500 2000 2500 3000
Time t [min]

Fig 8: Temporal variations in the surface tension values for Baker Petrolite studied at different
concentrations

5.3 Rotameter Calibration

Two rotameters have been employed in the present setup, one ranging from 0-3 LPM (low flowrates)
and the other ranging from 4-20 LPM (high flowrates). Calibration of these rotameters is done twice:
the first time is to check their performances relative to water (which is usually the standard calibrating
fluid). One their results are found satisfactory, the rotameters are then calibrated with the oil to be used
(Exxsol D80) to construct the equivalent calibration curve for the oil. These results are summarised
below:

Figs. 9(a),9(b): Calibration curves for the rotameters used in the Upflow Rig setup with respect to water

Following the calibration with water, the rotameters are calibrated with oil. Both rotameters are seen to
be consistent in recording flow measurement. A very minor effect, but something worth mentioning, is
that just like the O-rings, the surfactant also affects the rotameter float, part of which contains rubber.
Over time, the surfactant can be seen to chemically react with the rubber in the rotameter float; in fact
the effect is very much prominent to the naked eye.
40

Fig 10(a), 10(b): Comparison between the calibration curves for the rotameters used in the Upflow Rig setup
with respect to water and Exxsol D80

5.4 Experimental runs

The rotameters were found to be in good conditions and experimental runs were started. Upon
switching on the mass flowmeter, FlowDDE is executed to establish communication. This is then
followed by FlowPlot, to adjust all the flow parameters to “Default All” status, and the air line is then
opened. The gas flowrates are then then controlled electronically using the FlowView software. Once
the gas flowrate is steady in the test section, the centrifugal pump is switched on, enabling the oil Exxsol
D80 to enter the test section via the injector (in this case, an initial value of 500 LPM is given, that allows
the creation of a fairly stable liquid film). In each run, the liquid flowrate is set using the rotameter
pertaining to the desired flow. The flow is given some time to reach steady state, following which, the
Imperial i-speed 3 software is initiated, which triggers the high-speed camera. The camera records the
flow and produces a set of high-resolution images, which are then synthesised to produce videos
depicting the morphological behaviour of the system. Detailed flow insights and regime transitions are
obtained using this technique. Whenever a new surfactant concentration is to be used, the system is
first completely drained of oil, and left overnight. Upon filling the next day with oil, loaded with
surfactant Baker Petrolite, the centrifugal pump is primed, during equipment start-up to remove any
traces of air in the liquid line. Upon completion of runs, air is purged at very high flowrates (usually
around 1500-1800 LPM) to push any liquid that is present in the test section; this is especially important
as the surfactant has a very corrosive effect on Perspex. Higher surfactant concentration in the oil, does
lead to more crack formation in the test section.

Experimental data so recorded are then taken as inputs into a program coded in Dev C++, which then
predicts the fit of experimental data to the Wallis and Sun-Kutateladze correlations. Additionally, the
program also predicts the first onset of flooding for the oil-air system, when the Kutateladze number
attains a value of 3.2, which is the condition for flooding a proposed by Pushkina and Sorokin.
41

5.5 Experimental setup

The experimental setup as constructed for taking experimental runs may be represented as shown in
the schematic diagram below.

Fig 11: Schematic of the experimental setup used for experimental runs
42

6 Results and Discussions

The effects of surfactant are investigated on the interfacial dynamics, wave characteristics, regime
transitions, and flooding in vertical, counter-current, gas-liquid flows. High-speed axial imaging is
employed to understand regime behaviour, with Exxsol-D80 (oil) and Baker Petrolite (surfactant).
Surfactant-free runs are performed for comparison with data obtained for surfactant-laden systems
with different surfactant concentrations. Qualitative insight into possible candidates for exhibited
behaviour of flow regimes in the considered system.

6.1 Fitting of Experimental Data

The non-dimensional superficial liquid velocity as obtained from experimental runs was compared to
predicted non-dimensional liquid velocities obtained by simulations of Wallis and Sun-Kutateladze (which
are proposed for the air-water system).

Fig 12: Comparison of experimental results of the superficial liquid velocity with correlations proposed by Wallis
(R2=1) and Sun-Kutateladze(R2=0.885) as statistically estimated from SigmaPlot 13.0

Results indicate the Wallis correlation to be a better fit for the data, very possibly due to its more generic
definition. The correlations, though originally proposed for the air-water system, seem to agree with trends
seen in the air-oil system. At low values of the non-dimensional liquid superficial velocity, deviations
between these correlations are not markedly different from the experimental data. The deviations so seen
are in agreement with the investigative work of Clift et al. (1964) and the results of McQuillan and Whalley
(1985) that empirical correlations are seen to markedly deviate for systems involving more viscous fluids (as
compared to water). However, it is satisfying to see that the general trend of data obtained are similar across
both the multiphase flow systems.
43

6.2 Flooding profile: No surfactant

As the reference case, the flooding profile is tracked for the case where no surfactant is present in the system.
The flooding profile is as expected for a sharp edged entry case. A two-period, moving average, time-series
model is seen to be the best fit for this profile; similar conclusions have been drawn before by other
researchers (Dukler, 1984). The dotted orange line represents the time-series match to the experimental
profile, the latter is the orange line.

8.9
Gas Superficial Flooding Velocity----->

8.85

8.8

8.75

8.7

8.65

8.6

8.55

8.5

8.45

8.4
0 0.05 0.1 0.15 0.2 0.25 0.3
Liquid Superficial Velocity ------>

Fig 13: Flooding onset profiles without any surfactant loading, compared to a two-period, moving average time-series
model (best estimate)

6.3 Flooding profiles for different surfactant concentrations

As mentioned before, the onset of flooding is taken to be that point, when the climbing liquid film
successfully passes through the injector to the upper test section. For sharp edged entrances, the flooding
profile is essentially hyperbolic. The first estimate of the flooding onset is provided by the Pushkina-
Sorokin’s condition, which yields a value of 592 LPM (refer to Section 10.2). Actual experimental runs provide
a value of about 625 LPM, which decreases to about 620-618 LPM in the presence of surfactants. While the
44

addition of surfactant lowers the flooding point, this decrement is not very substantial.

Fig 14: Flooding profiles as a function of the non-dimensional, superficial velocities, plotted at different surfactant
loadings

The qualitative nature of these profiles follow the same trend as depicted in Fig 13. The behaviour of the air-
oil multiphase system follows the same trend as that of the air-water system. In each of these runs, the
flooding onset point is seen to drop down rapidly at lower liquid velocities, eventually plateauing off to a
more or less constant value, which remains unaffected with increase in the liquid flow rates. The present
experimental runs also confirm that Pushkina’s condition for the onset of flooding is valid approximately
not only to the air-water system, but also on the air-oil system. Further experiments may also be performed
to investigate if such a universal relationship may be devised that is valid for all such systems.

6.4 Interfacial Topology and Regime Transitions

High speed axial imaging is employed to capture images, which are then synthesised into videos using
VirtualDub, an open source video synthesis software by Microsoft. The relevant profiles are now shown
below. Each of these runs are taken at a fixed liquid volumetric flowrate (1.5 LPM) and for various gas
flowrates, characterised through the gas-phase Reynold’s number. From the images, it is but evident that the
surfactant Baker Petrolite has a “smoothening” effect on the interfacial topology. Additionally, wave
amplitude seems to decrease on increasing the surfactant loading; also there is a marked decrease in the
bubbles so formed with increase in the surfactant concentration. Some possible effects to explain these
observations would be tip-streaming and Marangoni stresses induced interface rigidification.
45

No surfactant (c=0 wppm):

4 4 4 4 4
Re =1.467*10 Re =2.293*10 Re =2.564*10 Re =2.768*10 Re =3.296*10
G G G G G

Downwards Flooding Bridging/Break Up Standing Waves Dry out

Low Surfactant concentration (c=500 wppm):

4 4 4 4 4
Re =1.467*10 Re =2.293*10 Re =2.564*10 Re =2.768*10 Re =3.296*10
G G G G G

Downwards Flooding/Bridging Standing Waves Dry out Dry out

High Surfactant concentration (c=1000 wppm):

4 4 4 4 4
Re =1.467*10 Re =2.293*10 Re =2.564*10 Re =2.768*10 Re =3.296*10
G G G G G

Downwards Flooding/Bridging Standing Waves Dry out Dry out

Fig 15: Qualitative high speed images of the interfacial topology under different experimentally tested flow conditions
46

Some qualitative insights may be immediately obtained from the above flow regime pictures. The surfactant
Baker Petrolite “smoothens” the flow, as compared to surfactant-free runs; flows with surfactant in oil have
lesser turbulent profiles. Addition of surfactant leads to interface rigidification, which means that surfactant
induced Marangoni stresses prevent interface deformations, which as a direct consequence, leads to reduced
wave amplitudes.

Increased surfactant concentration does not lead to any significantly sharper transitions between regimes,
however droplets in the flow are reduced. A possible candidate mechanism to explain this observation is tip-
streaming. The phenomena of tip streaming leads to the accumulation of surfactant at the extremities of a
sigmoidal shaped mother bubble, which then becomes very efficient in spitting out fine drops. The proposed
mechanisms are consistent with experimental findings.

6.5 Flow Smoothening, Drop and Bubble Reduction at Flooding

The surfactant Baker Petrolite tends to smoothen the flow, this effect is especially observable at the flooding
point for the system. Additionally, drops and bubbles are reduced at the flooding point, where the flow
transitions from downwards to annular, with increase in the surfactant concentration.

C=0 wppm C=500 wppm C=1000 wppm

Fig 16: Qualitative effects of surfactant addition on the air-oil multiphase system at the flooding point

At lower gas flowrates, the liquid flow is downwards and turbulent. Tip-streaming results in more
bubbles being formed of smaller radii, when surfactants are added. Increase in the surfactant
concentration would require a lower gas flowrate to achieve this effect; as a results most bubbles and
drops are dispelled prior to the flooding onset. This effect increases with increase in the surfactant
loading, as is evident from Fig 16. It has already been stated before that surfactant induced Marangoni
stresses lead to flow smoothening, as the interface become more resilient to shear stresses.
47

6.6 A Preliminary estimate to the Liquid Film Profile

Under very turbulent flows, the Wallis correlation is employed to obtain the increase in the liquid film
thickness by correlating it empirically to the non-dimensional liquid superficial velocity. This has been
modelled for the “no-surfactant” case to get a preliminary estimate of increases in the liquid film with
increased liquid loading. This formula is only used for a homogeneous liquid, free from any other mixed
components. Thus, the formula cannot be employed for cases involving surfactant in the oil.

Fig 17: Variation of the liquid film thickness with the liquid phase non-dimensional superficial velocity for the
“no-surfactant” case

The resulting profile is then subjected to preliminary statistical analysis and a log-normal distribution seen
to best fit the data. Elementary investigations on this distribution shows that the log-normal distribution
frequently arises in problems pertaining to geological phenomena, and also in the domain of fluid mechanics.
The next closest match for the produced profile is a Weibull distribution. Statistical analysis has been
performed on SigmaPlot 13.0. These results conclusively prove the validity and coherence between
experimental results and simulations, an ANOVA (Analysis of Variance) test proves this.
48

Nonlinear Regression

Data Source: Data 1 in Notebook1

Equation: Probability, Lognormal CDF


f=lognormdist(x,a,b)

R Rsqr Adj Rsqr Standard Error of Estimate


0.9998 0.9996 0.9996 8.6748E-006

Coefficient Std. Error t P


a 14.6676 0.1223 119.8831 <0.0001
b 5.1984 0.0393 132.1151 <0.0001

Analysis of Variance:
DF SS MS
Regression 2 6.0956E-006 3.0478E-006
Residual 11 8.2778E-010 7.5252E-011
Total 13 6.0964E-006 4.6895E-007
Corrected for the mean of the observations:
DF SS MS F P
Regression 1 2.1167E-006 2.1167E-006 28128.1933<0.0001
Residual 11 8.2778E-010 7.5252E-011
Total 12 2.1175E-006 1.7646E-007

Statistical Tests:
Normality Test (Shapiro-Wilk) Passed (P = 0.1704)
W Statistic= 0.9077 Significance Level = 0.0500

Constant Variance Test Passed (P = 0.6164)

Number of Iterations Performed = 15

Nonlinear Regression

Data Source: Data 1 in Notebook1

Equation: Probability, Weibull CDF


f=weibulldist(x,a,b)

R Rsqr Adj Rsqr Standard Error of Estimate


1.0000 1.0000 1.0000 2.4936E-007

Coefficient Std. Error t P


a 0.6674 0.0001 4564.9338 <0.0001
b 7818.2904 17.8412 438.2165 <0.0001

Analysis of Variance:
DF SS MS
Regression 2 6.0964E-006 3.0482E-006
Residual 11 6.8399E-013 6.2181E-014
Total 13 6.0964E-006 4.6895E-007
Corrected for the mean of the observations:
DF SS MS F P
Regression 1 2.1175E-006 2.1175E-006 34054743.3457<0.0001
Residual 11 6.8399E-013 6.2181E-014
Total 12 2.1175E-006 1.7646E-007

Statistical Tests:

Normality Test (Shapiro-Wilk) Passed (P = 0.7272)


W Statistic= 0.9583 Significance Level = 0.0500

Constant Variance Test Passed (P = 0.9494)

Number of Iterations Performed = 20

Fig 18: Statistical fitting of the film-thickness data on SigmaPlot 13.0


49

6.7 Discussions

The present research has demonstrated conclusively that surfactants can influence the onset of flooding in
vertical, counter-current, continuous-contact equipments. There are a couple of salient features that must be
noted.

The first conclusion that can be made is that the empirical correlations of researchers (Wallis, Sun and
Kutateladze etc.) which were originally formulated during investigations of the air-water system, are also
valid for the air-oil multiphase system. Values of these constants will obviously change to fit the new system
being studied, but the qualitative trends are the same. Additionally, the same trends are again followed when
surfactants are added to the system.

The addition of surfactant Baker Petrolite rapidly smoothens the flow, and this effect is drastically increased
with higher surfactant concentration. Surfactants also tend to reduce bubbles in the flow, and doubling the
concentration of surfactant from 500 wppm to 1000 wppm results in almost no bubbles in the flow. It is
customary to classify such vertical flows in terms of flow regimes and the surfactant markedly affects regime
transitions by lowering the gas velocity for the flooding onset, for a fixed liquid velocity. However, it is to be
noted that increased surfactant concentration lowers the flooding point only very slightly and therefore, its
increased use may be uneconomical, when seen from the industrial perspective. The flooding profile that
results is consistent with the trends predicted by Hewitt and Wallis, corresponding to a sharp edged gas
inlet.

The increase of the liquid film with increase in the liquid loading has been simulated using the empirical,
non-dimensional Wallis correlation. The profile is parabolic, and statistical curve fitting of the liquid film
profile indicates that a log-normal distribution is the best match for the data.

As a last mention, the surfactant Baker Petrolite,


though strong as other surfactants like Triton X-100, is
nevertheless corrosive on the Perspex tube. The above
diagram shows explicitly how the introduction of
surfactant on the upflow rig leads to the creation of
tiny, microstructural cracks within the tube walls. This
is a significant issue that needs to be tackled and found
a solution to in future.

Fig 19: Corrosive action of surfactant Baker Petrolite on the


Perspex tube of the rig
50

7 Conclusions

The effect of surfactants in altering the flow properties of the oil-air multiphase system has been investigated
in this present work. A clear understanding has emerged about how flow regimes are shifted upon
introducing surfactants to the studied system, and possible candidate mechanisms are proposed to explain
the flow behaviour.

At low gas velocities, the flow is turbulent and bubbles are seen in the test section. Increasing the surfactant
concentration leads to increase in the liquid entrainment and increases the bubbles in the test section. A
surfactant is a chemical that reduces the surface tension for the system. Owing to their molecular structure,
it is energetically favourable for them to reside at the free surface. Their tendency to reduce the surface
tension results in a special class of Marangoni flows. As a consequence, they tend to reduce the values of
normal and shear stresses for the system. When a bubble travels in a surfactant laden fluid, it experiences
less shear stress (so called resistance by the fluid), and hence, it is easier for the bubble to travel upward than
in the case with no surfactant. This effect is magnified with increase in the surfactant concentration, which is
why increasing the surfactant concentration results in more bubbles passing through the test section. It is to
be mentioned here that the effects pertaining to bubble increase are seen when the fluid is in the downward
flow regime.

With increase in the gas velocity, the flow passes onto the “flooding” regime, which is essentially a type of
annular-flow. In this regime, the onset of flooding is identified by the appearance of waves which eventually
carries the oil above the injector. As explained before, increase in the surfactant loading will smoothen the
flow, which reduces the wave amplitude. The candidate mechanism that can explain this behaviour is the
“interfacial levelling” effect that surfactants exhibit. The surfactant will tend to preferentially reside on the
gas-liquid interface, and tends to resist any interfacial deformations. Thus, when a wave propagates, these
surfactant induced Marangoni stresses will tend to resist the wave from attaining a large amplitude, as
otherwise this would result in interfacial disturbance. The Marangoni effect would therefore, lead to
suppression of surface waves. This so-called healing effect on the interface will lead to reductions in the
flooding wave amplitude, as is evident from the high-speed camera images. Additionally, as expected,
increase in the surfactant concentration would lead to this effect becoming more pronounced. One can see
in Fig 16 how ripple waves are smoothened out with increase in the surfactant loading. Also, further changes
in the surfactant loading beyond 500 wppm does not lead to a marked change in the flow. The hypothesis
proposed is consistent with the experimental findings of Duangprasert et al. (2010).

Droplets are also seen in the flow system. Droplets are more densely located at low gas flowrates, and tend
to lessen upon increasing the gas velocity. This is clearly evident by looking at the sequence of images
depicting the qualitative flow regime transitions for the no-surfactant case. An increase in the surfactant
concentration leads to a reduction in the droplet density for the same conditions of liquid and gas velocities.
51

The possible explanation of this may be the “tip-streaming” phenomena. A large sigmoid-shaped mother
drop usually tends to break into smaller daughter droplets; this effect is substantially enhanced upon
surfactant addition to the system. The surfactant tends to preferentially locate towards the droplet
extremities, and thus initiates the formation of more droplets through tip-streaming. The lessening of shear
stresses in the liquid aids the movement of these droplets, which continuously move upwards. Hence,
increasing the surfactant concentration would lead to droplets moving upwards more easily, therefore, for
the same flow conditions, lesser droplets reside on the test section. This effect happens rather drastically, and
increasing the surfactant loading from 500 wppm to 1000 wppm essentially results in almost complete
droplet removal from the test section, near the flooding point.

Bubbles are primarily seen in the downwards flow regime, just prior to the flooding point. With increase in
the surfactant concentration, more bubbles are seen to travel through the test section; however bubble radii
become smaller. In this case, the tip-streaming effect leads to the generation of very fine dispersed phases as
a corner or tip is created at the tip of the bubble, where the surfactant accumulates and hence, lowers the
surface tension. This makes the mother bubble very efficient in spitting out very fine bubbles. Increased
surfactant concentration would lead to this effect happening at lower gas flowrates, which is consistent with
the experimental findings of Vijayan et al. (2001).

These results are interesting when compared and contrasted with qualitative results of co-current, air-water
flows loaded with a surfactant of foaming nature (van Nimwegen, 2015). Unlike the present case where tip-
streaming and reduced shear stresses reduces and suppresses large droplet formation, foaming is proposed
as the major suppressing mechanism. Also, the researchers report distinct waves (roll waves, ripple waves,
foaming waves etc.) characterising different flow regimes, which is not seen for air-oil systems. A very
cursory comparison between the images captured proves this: while one can see distinct and different waves
in the air-water system, a predominantly similar ripple-type wave is seen across all flow regimes in case of
the air-oil system.

The surfactant has a corrosive effect on the Perspex tube, and repeated runs lead to the Perspex developing
microstructural fractures. This is a significant challenge that needs to be overcome, especially if SHELL seeks
to employ this technology in their equipments (for instance: their perdido risers). The interaction of the
surfactant with the tube material needs further investigations, prior to commercial deployment. Apart from
this, the next stage of this research should encompass a more quantitatively rigorous assessment of the flow
and extract some more useful parameters like the average film thickness, droplet entrainment rate, pressure
gradients etc.
52

8 Acknowledgement

I wish to thank Prof. Omar K. Matar for being my principal supervisor and mentor at Imperial College
London. Prof. Matar has been instrumental in helping me navigating through the present research, as well
as my time here at the department. He has always encouraged me, supported my academic and research
efforts and has always inspired me to perform to my very zenith. His subject knowledge and expertise on
fluid mechanics is truly outstanding and I consider it my privilege that despite his extremely busy schedule,
he always kept track of my research. Additionally, he has always given me candid, positive feedback which
I have utilised to elevate the quality of the present work. Prof. Matar has guided me in significant aspects of
my life, both professional and personal. I am indeed grateful to him for being there for me, whenever I
needed his experience.

Dr. Christos N. Markides, my co-supervisor and mentor, has encouraged and taught me how to take charge
and perform independent research. From designing and revamping the upflow rig almost single-handedly,
to performing experimental runs by myself and producing meaningful, insightful research, my postgraduate
thesis has taught me a lot, the credit for which goes to Dr. Markides. The skills I have learnt under him have
made me into a more pragmatic, hands-on chemical engineer. He has always been extremely supportive of
my efforts too, regularly following up on my progress and yet, giving me my own flexible time schedule to
plan the rig construction, leak fixing and operation. His efforts have made me cultivate a very diverse skillset,
which would not only put me at a competitive advantage in my field, but would also make me a better
researcher and an able chemical engineer. Dr. Markides has always given me constructive, positive feedback
whenever I have approached him, which is why even within such a limited time span, I was able to progress
the research forward substantially.

I also wish to thank the principal sponsor for this project, viz. SHELL UK Ltd. for allocating the necessary
funding to the project so as to enable research in this topic. I am convinced that this work will in turn help
the company in understanding the flow behaviour of the oil-air system, and just how surfactants have the
potential of altering the flow characteristics, more so at the onset of flooding. Such a detailed understanding
might potentially help the company on the economic frontier in the near future.
53

9 References

[1] Elgin J.C. and Weiss F.B. Liquid Holdup and Flooding in Packed Towers. Ind. Eng. Chem. 1939; 31(4):

435-445.

[2] Sherwood T.K, Shipley G.H. and Holloway F.A.L. Flooding Velocities in Packed Columns. Ind. Eng.

Chem. 1938; 30(7):765-769.

[3] Uchida, S., and Fujita, S. Irrigated Packed Towers. J. Soc. Chem. Ind. (Japan); 1939: 886 (1936).

[4] Cetinbudaklar A.G and Jameson G.J. The mechanism of flooding in vertical countercurrent two-phase flow.

Chem. Eng. Science 1969; 24: 1669-1680.

[5] Clift R., Pritchard C.L. and Nedderman R.M. The effect of viscosity on the flooding conditions in wetted

wall columns. Chem. Eng. Science 1966; 21: 87-95.

[6] Cohen L.S. and Hanratty T.J. Effect of waves at a gas-liquid interface on a turbulent air flow. J. Fluid Mech

1968; 31: 467-479.

[7] Kamei S., Oishi J. and Okane T. Flooding in a Wetted Wall Tower. Chem Engng. (Jap) 1954; 18: 364-368.

[8] Suzuki S. and Ueda T. Behaviour of Liquid Films And Flooding In Counter-current Two Phase Flow-Part

I. Flow in Circular Tubes. Int. J. Multiphase Flow 1977; 3: 517-532.

[9] Kusuda H. and Imura H. Stability of a liquid film in a counter-current annular two-phase flow. Trans. Jap.

Soc. Mech. Engrs 1974; 40: 1082-1088.

[10] Levich V. Physicochemical Hydrodynamics 1962; 620. Prentice-Hall, New York.

[11] Portalski S. and Cleeg A.J. An experimental study of wave inception on falling liquid films. Chem.

Engng Sci. 1972; 27: 1257-1265.

[12] Wallis G.B and Makkenchery S. The Hanging Film Phenomenon in Vertical Annular Two-Phase Flow.

ASME J. Fluids Engng. 1974; 96: 297-298.

[13] McQuillan K.W and Whalley P.B. A Comparison Between Flooding Correlations And Experimental

Flooding Data For Gas-Liquid Flow in Vertical Circular Tubes. Chem. Engg. Science 1985; 40 (8): 1425-

1440.

[14] Pushkina O.L and Sorokin Y.L. Breakdown of Liquid Film Motion in Vertical Tubes. Heat Transfer-Soviet

Research 1969; 1 (5): 56-64.


54

[15] Maron D.M and Dukler A.E. Flooding and Upward Film Flows in Vertical Tubes - II. Speculations on

Film Flow Mechanisms. Int. J. Multiphase Flow 1984; 10 (5): 599-621.

[16] Hewitt G.F and Wallis G.B. Flooding and associated phenomena in falling film flow in a vertical tube.

UKAEA Report, 1961 AERE R-4022.

[17] Vijayan M, Jayanti S and Balakrishnan A.R. Effect of Tube Diameter on Flooding. Int. J. Multiphase

Flow 2001; 27: 797-816.

[18] Govan A.H, Hewitt G.F, Richter H.J, Scott A. Flooding and Churn Flow in Vertical Pipes. Int. J.

Multiphase Flow 1991; 17: 27-44.

[19] Zabaras G.J and Dukler A.E. Countercurrent gas-liquid annular flow, including the flooding state.

AIChE 1988; 34: 389-396.

[20] Biage M. 1989. Structure de la surface libre d'un film liquide ruisselant sur une plaque plane verticale et

soumis a un contrecourant de gas: transition vers l'ecoulement cocourant ascendant. Ph.D. thesis,

Institut National Polythenique de Grenoble, France.

[21] Dukler A.E, Smith L and Chopra A. Flooding and Upward Film Flows in Tubes -I. Experimental

Studies. Int. J. Multiphase Flow 1984; 10(5): 585-597.

[22] Wallis G.B. Flooding Velocities for Air and Water in Vertical Tubes 1961. AAEW-R123, UKAEA,

Harwell, England.

[23] Hewitt G.F and Wallis G.B. Flooding and associated phenomena in falling film flow in a vertical tube.

1963 AAEW-R4022, UKAEA, England.

[24] Sun K.H. Proceedings of Multiphase Flow and Heat Transfer Symposium Workshop. Miami beach,

Florida 1979: 1615-1635.

[25] Alekseev V.P, Poberezkin A.E and Gerasimov P.V. Determination of Flooding ranges in regular

packings. Heat Trans-Soviet Res 1972; 4: 159-163.

[26] Jayanti S, Tokarz A and Hewitt G.F. Theoretical investigation of the diameter effect on flooding in

countercurrent flow. Int. J. Multiphase Flow 1996; 22: 307-324.

[27] Hewitt G.F, Lacey P.M.C and Nicholls B. Transition in film flow in a vertical flow. 1965 AERE-

R4614. UKAEA, Harwell, England.

[28] Jeong J.H and No H.C. Experimental Study of the Effect of Pipe Length and Pipe-End Geometry on
55

Flooding. Int. J. Multiphase Flow 1996; 22(3): 499-514.

[29] Bankoff S.G and Lee S.C. Critical Review of the Flooding Literature. Multiphase Science and Technology

1986; 2: 95-180.

[30] Deendarlianto , Ousaka A, Indarto, Kariyasaki A, Lucas D, Vierow K, Vallee C and Hogan K. The

effects of surface tension on flooding in counter-current two-phase flow in an inclined tube. Experimental

Thermal and Fluid Science 2010; 34: 813-826.

[31] Modern Chemical Enhanced Oil Recovery. Elsevier Press. DOI: 10.1016/B978-1-85617-745-0.00007-3.

[32] Farn R.J., ed. Chemistry and technology of surfactants. Oxford: Blackwell Publishing, 2006.

[33] Liu L, Li X, Tong L and Liu Y. Effect of surfactant additive on vertical two phase flow. Journal of

Petroleum Science and Engineering 2014. http://dx.doi.org/10.1016/j.petrol.2014.02.004

[34] Rozenblit R, Gurevich M, Lengel Y and Hetsroni G. Flow patterns and heat transfer in vertical

upward air-water flow with surfactant. Int J. Multiphase Flow 2006; 32: 889-901.

[35] Guo-dong X and and Lei C. Influence of surfactant on two-phase flow regime and pressure drop

in upward inclined pipes. J. of Hydrodynamics 2012; 24(1): 39-49.

[36] van Nimwegen A.T, Portela L.M and Henkes R.A.W.M. The Effect of Surfactants on Vertical

Air/Water Flow for Prevention of Liquid Loading. SPE 164095. From the SPE International

Symposium on Oilfield Chemistry in The Woodlands, Texas, USA. 8-10 April, 2013.

[37] van Nimwegen A.T, Portela L.M and Henkes R.A.W.M. The effect of surfactants on air-water annular

and churn flow in vertical pipes. Part I: Morphology of the air-water interface. Int. J. Multiphase Flow

2015; 71: 133-145.

[38] Taitel Y, Bornea D and Dukler A.E. Modelling Flow Pattern Transitions for Steady Upward Gas-Liquid

Flow in Vertical Tubes. AIChE Journal May 1980; 26(3): 345-354.

[39] Zargartalebi M, Kharrat R and Barati N. Enhancement of surfactant flooding performance by the use of

silica nanoparticles. Fuel 2015; 143: 21-27.

[40] Khosla V. Visual Investigation of Annular Flow and the Effect of Wall Wettability. MSc. thesis, Delft

University of Technology, 2012.

[41] Boos J, Drenckhan W and Stubenrauch C. On How Surfactant Depletion during Foam Generation

Influences Foam Properties. Langmuir (ACS Publication) 2012; 28: 9303-9310.

[42] Mao Z.S and Dukler A.E. The myth of churn flow? Int. J. Multiphase Flow 1993; 19: 377-383.
56

[43] Hewitt G.F and Jayanti S. To Churn or Not to Churn. Int J. Multiphase Flow 1993; 19(3): 527-529.

[44] Bharathan D and Wallis G.B. Air-Water Countercurrent Annular Flow. Int. J. Multiphase Flow 1983;

9(4): 349-366.

[45] Wallis G.B. One Dimensional Two-Phase Flow. p. 336. McGraw Hill, New York.

[46] Anderson G.H and Mantzouranis B.G. Two phase (gas-liquid) flow phenomena-I Pressure drop and

holdup for two-phase flow in vertical tubes. Chemical Engineering Science 1960; 12: 109-126.

[47] Jayanti S, Hewitt G.F, Low D.E.F and Herveiu E. Observation of Flooding in The Taylor Bubble of Co

Current Upwards Slug Flow. Int. J. Multiphase Flow 1993; 19(3): 531-534.

[48] Vijayan M, Jayanti S and Balakrishnan A.R. Experimental study of air-water countercurrent annular

flow under post-flooding conditions. Int. J. Multiphase Flow 2002; 28: 51-67.

[49] Feind F. Falling liquid films with countercurrent air flow in vertical tubes VDI-Forschungsh 1960; 26:

481.

[50] Zapke A and Kroger D.G. Countercurrent Gas-Liquid flow in inclined and vertical ducts-I : Flow

patterns, pressure drop characteristics and flooding. Int. J. Multiphase Flow 2000; 26: 1439-1455.

[51] Zapke A and Kroger D.G. Countercurrent Gas-Liquid flow in inclined and vertical ducts-II : The

validity of the Froude-Ohnesorge number correlation for flooding. Int. J. Multiphase Flow 2000; 26: 1457-

1468.

[52] Chung K.S, Liu C.P and Tien C.L. Flooding in two-phase countercurrent flows-I: Experimental

investigation. PhysicoChemical Hydrodynamics 1980; 1: 195-207.

[53] Chung K.S, Liu C.P and Tien C.L. Flooding in two-phase countercurrent flows-II: Experimental

investigation. PhysicoChemical Hydrodynamics 1980; 1: 209-220.

[54] Govan A.H, Hewitt G.F, Owen D.G and Burnett G. Wall Shear Stress Measurements in Vertical

Air-Water Annular Two-Phase Flow. Int. J. Multiphase Flow 1989; 15(3): 307-325.

[55] Govan A.H, Hewitt G.F and Lim S.B. Visualisation Studies of Churn Flow. Institution of Chemical

Engineers, Second Irish Research Colloquia. Belfast, 9 March 1990.

[56] Govan A.H, Hewitt G.F, Richter J and Scott A. Flooding and Churn Flow in Vertical Pipes. Int. J.

Multiphase Flow 1991; 17(1): 27-44.


57

[57] Barbosa Jr. J.R, Govan A.H and Hewitt G.F. Visualisation and Modelling Studies of churn flow in a

vertical pipe. Int. J. Multiphase Flow 2001; 27: 2105-2127.

[58] Wolf A, Jayanti S and Hewitt G.F. Flow development in vertical annular flow. Chemical Engineering

Science 2001; 56: 3221-3235.

[59] Taitel Y and Dukler A.E. A Model for Slug Frequency during Gas-Liquid Flow in Horizontal and Near

Horizontal Pipes. Int. J. Multiphase Flow 1977; 3: 585-596.

[60] Taitel Y and Dukler A.E. A theoretical approach to the Lockhart-Martinelli correlation for stratified flow.

Int. J. Multiphase Flow 1976; 2: 591-595.

[61] Lockhart R.W and Martinelli R.C. Proposed correlation of data for isothermal two-phase, two-component

flow in pipes. Chem. Engng Prog. 1949; 45: 39-48.

[62] Johannessen T. A theoretical solution for the Lockhart and Martinelli flow model for calculating two

phase flow pressure drop and holdup. Int. J. Heat Mass Transfer 1972; 15: 1443-1449.

[63] Lighthill M.H and Whitham G.F. On kinematic waves. Proc. R. Soc. 1955; 229: 281-344.

[64] Taitel Y and Barnea D. A Film Model for the prediction of Flooding and Flow Reversal for Gas-Liquid

Flow in Vertical Tubes. Int. J. Multiphase Flow 1982; 8(1): 1-10.

[65] Lee S.C and Bankoff S.G. Parametric effects on the onset of flooding in flat-plate geometries. Int. J. Heat

Mass Transfer 1984; 27(10): 1691-1700.

[66] Ousaka A, Deendarlianto, Kariyasaki A and Fukano T. Prediction of flooding gas velocity in gas

liquid counter-current two-phase flow in inclined pipes. Nuclear Engineering and Design 2006; 236:

1282-1292.

[67] Dukler A.E. A Physical Model for Predicting the Minimum Stable Slug Length. Chemical Engg. Science

1985; 40(8): 1379-1385.

[68] Salazar R.P and Marschall E. Time-average local thickness measurement in falling liquid film flow. Int.

J. Multiphase Flow 1978; 4: 405-412.

[69] Lacy C.E and Dukler A.E. Flooding in Vertical Tubes-I Experimental Studies for the Entry Region. Int.

J. Multiphase Flow 1994; 20(2): 219-233.

[70] Lacy C.E. and Dukler A.E. Flooding in Vertical Tubes-II A Film Model for Entry Region flooding. Int.

J. Multiphase Flow 1994; 20(2): 235-247.


58

[71] Paglianti A, Pintus S and Giona M. Time-series analysis for the identification of flooding/loading

transition in gas-liquid stirred tank reactors. Chemical Engg. Science 2000; 55: 5793-5802.

[72] Hlaing N.D, Sirivat A, Siemanond K and Wilkes J.O. Vertical two-phase flow regimes and pressure

gradients: Effect of Viscosity. Experimental Thermal and Fluid Science 2007; 31: 567-577.

[73] Nicklins D.J, Wilkes J.O and Davidson J.F. Two-phase flow in vertical tubes, Transactions of the

Institute of Chemical Engineers 1962; 40(3): 61-68.

[74] Deendarlianto, Ousaka A, Kariyasaki A and Fukano T. Investigation of liquid film behaviour at the

onset of flooding during adiabatic counter-current air-water two-phase flow in an inclined pipe. Nuclear

Engg. and Design 2005; 235: 2281-2294.

[75] Karimi G and Kawaji M. Flooding in vertical counter-current annular flow. Nuclear Engg. and Design

2000; 200: 95-105.

[76] Waltrich P.J, Falcone G and Barbosa Jr. J.R. Liquid transport during gas flow transients applied to

liquid loading in long, vertical pipes. Experimental Thermal and Fluid Science 2015; 68: 652-662.

[77] Johnston A.J. An investigation into the interfacial shear stress contribution in two-phase stratified flow.

Int. J. Multiphase Flow 1984; 10(3): 371-383.

[78] Guedes de Carvalho J.R.F and Talaia M.A.R. Interfacial shear stress as a criterion for flooding in

counter current film flow along vertical surfaces. Chemical Engg. Science 1998; 53(11): 2041-2051.

[79] Kirillov P.L, Kashcheyev V.M, Muranov Yu. V and Yuriev Yu. S. A two-dimensional mathematical

model of annular-dispersed and dispersed flows-I. Int. J. Heat Mass Transfer 1987; 30(4): 791-800.

[80] Choutapalli I and Vierow K. Wall pressure measurements of flooding in vertical countercurrent annular

air-water flow. Nuclear Engg. and Design 2010; 240: 3221-3230.

[81] Bezrodnyj M.K and Antoshko Yu.V. Critical Regimes of Upward Film Flows in Vertical annular

Channels. Experimental Thermal and Fluid Science 1992; 5: 448-456.

[82] Whalley P.B. Brief communication: Flooding, Slugging and Bottle Emptying. Int. J. Multiphase Flow

1987; 13(5): 723-728.

[83] Richter H.J. Flooding in Tubes and Annuli. Int. J. Multiphase Flow 1981; 7(6): 647-658.

[84] Cioncolini A and Thome J.R. Prediction of the entrained liquid fraction in vertical annular gas-liquid
59

two-phase flow. Int. J. Multiphase Flow 2010; 36: 293-302.

[85] Drosos E.I.P, Paras S.V and Karabelas A.J. Counter-current gas-liquid flow in a vertical narrow

channel- Liquid film characteristics and flooding phenomena. Int. J. Multiphase Flow 2006; 32: 51-81.

[86] Nicklin D.J and Davidson J.F. The onset of instability in two-phase slug flow. Institution of Mech. Engrs,

Proc. Symposium on Two-Phase Flow, 7th February 1962.

[87] Davidson, J.F and Shearer C.J. Flooding in wetted wall columns: its relation to the formation of a standing

wave due to gas blowing upwards over a vertical liquid film. J. of Fluid Mech. 1965; 321-335.

[88] Schutt J.B. A theoretical study of the phenomenon of bridging in wetted-wall columns. PhD. Thesis,

University of Rochester, New York 1959.

[89] Chakraborty S. The Effect of Surfactant on Flooding Phenomena in Vertical, Counter-current, Gas-Liquid

Flows (part of a SHELL UK funded project). Literature Review submitted to Imperial College London

2016.
60

10 Appendix

10.1 Program in Dev C++ Ver. 5.4.2 for data extraction

#include<stdio.h>
#include<conio.h>
#include<math.h>

int main() // Program begins. Mainclass starts from here!


{

// STEP 1:(GENERAL DESCRIPTION)General info- MSc. thesis flooding calculations


printf("\n Coded by: Sourojeet Chakraborty, Imperial College London.");
printf("\n SHELL rig flooding calculations begin now ");

// STEP 2: Variable Declaration and Initialization


float L=0;//Length of the test section.
float D=0;//Diameter of the test section.
float sigma=0;//Surface Tension for ONLY oil i.e Exxsol D-80
float rho_l=0;
float rho_g=0;
float UgP=0;
int nl=0;
int nv=0;
int i=0;

printf("\n Enter the number of data points for the liquid flow rate: ");
scanf ("%d",&nl);
printf("\n Enter the number of data points for the vapour flow rate: ");
scanf ("%d",&nv);
printf("\n Enter the test section length <m>: "); //L=1.44m
scanf ("%f",&L);
printf("\n Enter the test section diameter <m>: "); //D=0.04m
scanf ("%f",&D);
printf("\n Enter the surface tension <N/m>: "); //sigma=0.0263 N/m
scanf ("%f",&sigma);
printf("\n Enter the liquid density <kg/m3>: ");
scanf ("%f",&rho_l);
printf("\n Enter the vapour density <kg/m3>: ");
scanf ("%f",&rho_g);

// STEP 3: Declaration and Initialization of Arrays


float Ql[nl]; //Flowrates.
float Qg[nv];
float Ul[nl]; //Superficial Velocities.
float Ug[nv];
float Ul_star[nl]; //Non-Dimensional Velocities.
float Ul_star2[nl];
float Ul_star4[nl];
float Ug_star[nv];
float Ug_star2[nv];
float Ug_star4[nv];

float KuL_star[nl];
61

float KuL_star4[nl];
float KuG_star[nv];
float KuG_star4[nv];

float delta[nl]; //Flim Thickness Matrix

for(i=0;i<nl;i++)
{
Ql[i]=0;
Ul[i]=0;
Ul_star[i]=0;
Ul_star2[i]=0;
Ul_star4[i]=0;
KuL_star[i]=0;
KuL_star4[i]=0;
delta[i]=0;
}
for(i=0;i<nv;i++)
{
Qg[i]=0;
Ug[i]=0;
Ug_star[i]=0;
Ug_star2[i]=0;
Ug_star4[i]=0;
KuG_star[i]=0;
KuG_star4[i]=0;
}
for(i=0;i<nl;i++)
{
printf("\n Enter the liquid flowrate <LPM>: ");
scanf ("%f",&Ql[i]);
}

for(i=0;i<nv;i++)
{
printf("\n Enter the vapour flowrate <LPM>: ");
scanf ("%f",&Qg[i]);
}

//STEP 4: Evaluating the superficial and non-dimensional velocities.


for(i=0;i<nl;i++)
{
Ul[i]=(0.000017*Ql[i])/(0.25*3.1416*D*D);
Ul_star[i]=Ul[i]*pow((rho_l/(9.81*D*(rho_l - rho_g))),0.5);
}

for(i=0;i<nv;i++)
{
Ug[i]=(0.000017*Qg[i])/(0.25*3.1416*D*D);
Ug_star[i]=Ul[i]*pow((rho_g/(9.81*D*(rho_l - rho_g))),0.5);
}

//STEP 5: Printing the velocities


printf("\n Superficial Liquid velocities <m/s>: ");
for(i=0;i<nl;i++)
62

printf("\n %f",Ul[i]);

printf("\n Non-Dimensional Superficial Liquid velocities: ");


for(i=0;i<nl;i++)
printf("\n %f",Ul_star[i]);

printf("\n Superficial Vapour velocities <m/s>: ");


for(i=0;i<nv;i++)
printf("\n %f",Ug[i]);

printf("\n Non-Dimensional Superficial Vapour velocities: ");


for(i=0;i<nv;i++)
printf("\n %f",Ug_star[i]);

// STEP 6: Obtaining and printing the square and fourth roots of the non-dimensional superficial
velocities
for(i=0;i<nl;i++)
{
Ul_star2[i]=pow(Ul_star[i],0.5);
Ul_star4[i]=pow(Ul_star[i],0.25);
}
for(i=0;i<nv;i++)
{
Ug_star2[i]=pow(Ug_star[i],0.5);
Ug_star4[i]=pow(Ug_star[i],0.25);
}
printf("\n Non-Dimensional Square Root of Liquid Superficial Velocities: ");
for(i=0;i<nl;i++)
printf("\n %f",Ul_star2[i]);

printf("\n Non-Dimensional Fourth Root of Liquid Superficial Velocities: ");


for(i=0;i<nl;i++)
printf("\n %f",Ul_star4[i]);

printf("\n Non-Dimensional Square Root of Vapour Superficial Velocities: ");


for(i=0;i<nv;i++)
printf("\n %f",Ug_star2[i]);

printf("\n Non-Dimensional Fourth Root of Vapour Superficial Velocities: ");


for(i=0;i<nv;i++)
printf("\n %f",Ug_star4[i]);

// STEP 7: Pushkina-Sorokin's equation for the predicted flooding value


UgP=3.2*(9.81*sigma*(rho_l - rho_g)/pow(rho_g,0.5));
printf ("\n Pushkina-Sorokin's predicted value of flooding is :");
printf ("\n %f",UgP);

// STEP 8: Sun-Kutateladze Correlation Calculations


for(i=0;i<nl;i++)
{
KuL_star[i]=(Ul[i]*pow(rho_l,0.5))/pow((9.81*sigma*(rho_l - rho_g)),0.25);
KuL_star4[i]=pow(KuL_star[i],0.25);
}
for(i=0;i<nv;i++)
{
63

KuG_star[i]=(Ug[i]*pow(rho_g,0.5))/pow((9.81*sigma*(rho_l - rho_g)),0.25);
KuG_star4[i]=pow(KuL_star[i],0.25);
}
printf("\n Non-Dimensional Liquid Kutateladze Number: ");
for(i=0;i<nl;i++)
printf("\n %f",KuL_star[i]);

printf("\n Non-Dimensional Fourth Root of Liquid Kutateladze Number: ");


for(i=0;i<nl;i++)
printf("\n %f",KuL_star4[i]);

printf("\n Non-Dimensional Vapour Kutateladze Number: ");


for(i=0;i<nv;i++)
printf("\n %f",KuG_star[i]);

printf("\n Non-Dimensional Fourth Root of Vapour Kutateladze Number: ");


for(i=0;i<nv;i++)
printf("\n %f",KuG_star4[i]);

// STEP 9: Evaluation of the film thickness by the Wallis-Correlation


for(i=0;i<nl;i++)
{
delta[i]=D*0.063*pow(Ug_star[i],0.667);
}
printf("\n The film thickness using the Wallis correlation is: ");
for(i=0;i<nl;i++)
printf("\n %f",delta[i]);

getch();
}// End of Program.

10.2 Sample Output for the program presented in 10.1


Coded by: Sourojeet Chakraborty, Imperial College London.
SHELL rig flooding calculations begin now
Enter the number of data points for the liquid flow rate: 13

Enter the number of data points for the vapour flow rate: 13

Enter the test section length <m>: 1.44

Enter the test section diameter <m>: 0.04

Enter the surface tension <N/m>: 0.0263

Enter the liquid density <kg/m3>: 796

Enter the vapour density <kg/m3>: 1.225

Enter the liquid flowrate <LPM>: 0.3


64

Enter the liquid flowrate <LPM>: 0.5

Enter the liquid flowrate <LPM>: 1

Enter the liquid flowrate <LPM>: 1.5

Enter the liquid flowrate <LPM>: 2

Enter the liquid flowrate <LPM>: 2.5

Enter the liquid flowrate <LPM>: 3

Enter the liquid flowrate <LPM>: 4

Enter the liquid flowrate <LPM>: 6

Enter the liquid flowrate <LPM>: 8

Enter the liquid flowrate <LPM>: 10

Enter the liquid flowrate <LPM>: 15

Enter the liquid flowrate <LPM>: 20

Enter the vapour flowrate <LPM>: 655

Enter the vapour flowrate <LPM>: 640

Enter the vapour flowrate <LPM>: 638

Enter the vapour flowrate <LPM>: 635

Enter the vapour flowrate <LPM>: 628

Enter the vapour flowrate <LPM>: 625

Enter the vapour flowrate <LPM>: 624

Enter the vapour flowrate <LPM>: 625

Enter the vapour flowrate <LPM>: 625

Enter the vapour flowrate <LPM>: 625

Enter the vapour flowrate <LPM>: 624

Enter the vapour flowrate <LPM>: 625

Enter the vapour flowrate <LPM>: 625

Superficial Liquid velocities <m/s>:

0.004058
0.006764
65

0.013528
0.020292
0.027056
0.033820
0.040584
0.054113
0.081169
0.108225
0.135281
0.202922
0.270563

Non-Dimensional Superficial Liquid velocities:

0.006484
0.010806
0.021613
0.032419
0.043225
0.054032
0.064838
0.086451
0.129676
0.172901
0.216126
0.324190
0.432253

Superficial Vapour velocities <m/s>:

8.860931
8.658009
8.630953
8.590368
8.495671
8.455087
8.441559
8.455087
8.455087
8.455087
8.441559
8.455087
8.455087

Non-Dimensional Superficial Vapour velocities:

0.000254
0.000424
0.000848
0.001272
0.001696
0.002120
0.002544
0.003391
0.005087
66

0.006783
0.008479
0.012718
0.016957

Non-Dimensional Square Root of Liquid Superficial Velocities:

0.080522
0.103953
0.147012
0.180053
0.207907
0.232447
0.254633
0.294025
0.360105
0.415814
0.464894
0.569377
0.657459

Non-Dimensional Fourth Root of Liquid Superficial Velocities:

0.283764
0.322418
0.383422
0.424326
0.455968
0.482128
0.504612
0.542241
0.600088
0.644836
0.681831
0.754570
0.810839
Non-Dimensional Square Root of Vapour Superficial Velocities:

0.015949
0.020589
0.029118
0.035662
0.041179
0.046039
0.050434
0.058236
0.071324
0.082358
0.092079
0.112773
0.130219

Non-Dimensional Fourth Root of Vapour Superficial Velocities:

0.126287
67

0.143490
0.170640
0.188844
0.202926
0.214568
0.224574
0.241321
0.267065
0.286980
0.303445
0.335817
0.360859

Pushkina-Sorokin's predicted value of flooding is:


592.858276

Non-Dimensional Liquid Kutateladze Number:

0.030259
0.050431
0.100862
0.151293
0.201724
0.252155
0.302586
0.403448
0.605172
0.806896
1.008619
1.512929
2.017239

Non-Dimensional Fourth Root of Liquid Kutateladze Number:

0.417073
0.473887
0.563549
0.623670
0.670177
0.708626
0.741672
0.796979
0.882002
0.947773
1.002148
1.109059
1.191761

Non-Dimensional Vapour Kutateladze Number:

2.591673
2.532321
2.524408
2.512538
2.484840
68

2.472970
2.469013
2.472970
2.472970
2.472970
2.469013
2.472970
2.472970

Non-Dimensional Fourth Root of Vapour Kutateladze Number:

0.417073
0.473887
0.563549
0.623670
0.670177
0.708626
0.741672
0.796979
0.882002
0.947773
1.002148
1.109059
1.191761

The film thickness using the Wallis correlation is:

0.000010
0.000014
0.000023
0.000030
0.000036
0.000041
0.000047
0.000057
0.000074
0.000090
0.000105
0.000137
0.000166

View publication stats

You might also like